{ "0910/0910.0878_arXiv.txt": { "abstract": "We compare the accuracy of various methods for determining the transfer of the diffuse Lyman continuum in H{\\sc\\,ii} regions, by comparing them with a high-resolution discrete-ordinate integration. We use these results to suggest how, in multidimensional dynamical simulations, the diffuse field may be treated with acceptable accuracy without requiring detailed transport solutions. The angular distribution of the diffuse field derived from the numerical integration provides insight into the likely effects of the diffuse field for various material distributions. ", "introduction": "In this paper, we model the diffuse field structure of astrophysical H{\\sc\\,ii} regions. Recombinations within these regions emit radiation in the various hydrogen line spectra and Balmer and higher continua which is an observed characteristic of them. Emission in the Lyman continuum, however, is energetic enough to ionize other hydrogen atoms, and so is trapped within the nebula. This radiation field is believed to have a significant fraction of the energy density of the direct ionizing continuum in some parts of the H{\\sc\\,ii} region, and so it is important to model it correctly. This is particularly relevant when modelling the complex dynamics of the nebulae, or the emission from the internal features such as the tails of cometary globules in the Helix planetary nebula \\citep{odell07}. \\cite{ritze05} has suggested that diffuse fields may be particularly important at the edges of H{\\sc\\,ii} regions, in some regimes. This would seem to suggest that the diffuse field may have a more significant impact on their overall evolution than previously assumed. However, Ritzerveld uses a simple outward-only treatment of the diffuse radiation transport, and also assumes that the absorption coefficient is comparable for diffuse and direct radiation fields, although it is acknowledged that in reality the direct photons are likely to have a harder spectrum and will thus be significantly more penetrating. To investigate the validity of these conclusions, in this paper we apply detailed discrete-ordinate angular integration to investigate the validity of several approximate numerical transport schemes. To simplify the problem, we assume a pure-hydrogen nebula, without dust, and use a simple two-frequency approximation to the radiation flow. With our more detailed modelling, we find that the diffuse field can indeed dominate for the situations Ritzerveld describes. However, where the diffuse field dominates, it will typically also be outwardly-beamed and therefore be indistinguishable from the direct field for most purposes. For most astrophysically relevant conditions, the usual on-the-spot approximation is shown to give accurate results through most of a spherical nebula, except for a region close to the star where it {\\it overestimates}\\/ the diffuse field. We also assess the accuracy of the Eddington diffusion approximation for the diffuse field transfer, which may be a useful means of modelling diffuse transport effects in multidimensional simulations. In the present paper, we neglect the effects of dust and heavy elements in the H{\\sc\\,ii} region. This is a reasonable assumption for the case of cosmological H{\\sc\\,ii} regions; however, there is observational evidence for the importance of dust absorption within the ionized gas in galactic H{\\sc\\,ii} regions \\citep{cesar00,robbe05}. We briefly discuss the impact dust might have on our results. ", "conclusions": "We present detailed calculations of radiation transfer within a photoionized nebula, for a simplified physical model. We find that the diffuse field is indeed enhanced in the outer parts of such a nebula if there are steep density gradients in the H{\\sc\\,ii} region, particularly if the absorption cross section for diffuse and direct photons is comparable. These circumstances are not likely to be typical. We also find that when the diffuse radiation is relatively strong, it is also strongly beamed in the radial direction, and so the dynamical effect of the radiation field will in any case be similar to the direct illumination. We discuss the possibility of using a diffusion approximation for the diffuse radiation field in two- and three-dimensional radiation hydrodynamic calculations of H{\\sc\\,ii} regions. This may allow the diffuse field effects to be calculated to reasonable accuracy without requiring a full radiation transfer calculation. This paper has not considered in detail a number of additional processes which may effect the radiation field within the nebulae, such as the spectral hardening of radiation close to the ionization front, dust absorption and scattering, and the effects of helium and other heavy elements. This is left for future work." }, "0910/0910.2799_arXiv.txt": { "abstract": "{We report the results of a high-energy multi-instrumental campaign with \\integ, \\xte, and \\swift\\ of the recently discovered \\integ\\ source \\igr. The \\swift/XRT data allow us to refine the position of the source to RA$_{\\mathrm{J}2000}$= 19$^h$ 29$^m$ 55.9$^s$ Dec$_{\\mathrm{J}2000}$=+18$^\\circ$ 18\\arcmin\\ 38.4\\arcsec\\ ($\\pm 3.5$\\arcsec), which in turn permits us to identify a candidate infrared counterpart. The \\swift\\ and \\xte\\ spectra are well fitted with absorbed power laws with hard ($\\Gamma \\sim 1$) photon indices. During the longest \\swift\\ observation, we obtained evidence of absorption in true excess to the Galactic value, which may indicate some intrinsic absorption in this source. We detected a strong (P=40\\%) pulsation at $12.43781 (\\pm0.00003)$~s that we interpret as the spin period of a pulsar. All these results, coupled with the possible 117 day orbital period, point to \\igr\\ being an HMXB with a Be companion star. However, while the long-term \\integ/IBIS/ISGRI 18--40 keV light curve shows that the source spends most of its time in an undetectable state, we detect occurrences of short ($\\sim 2000-3000$~s) and intense flares that are more typical of supergiant fast X-ray transients. We therefore cannot make firm conclusions on the type of system, and we discuss the possible implications of \\igr\\ being an SFXT. } ", "introduction": "\\indent The INTErnational Gamma-Ray Astrophysics Laboratory (\\integ) has permitted a large number of new sources to be discovered. Amongst the $\\sim250$ new sources\\footnote{see http://isdc.unige.ch/$\\sim$rodrigue/html/igrsources.html for an up to date list of all sources and their properties.}, \\integ\\ has unveiled two new or poorly-known types of high mass X-ray binaries (HMXB): the very absorbed supergiant HMXBs, and the supergiant fast X-ray transients (SFXT). While all types of HMXBs (including those hosting a Be star, Be-HMXBs) are powered by accretion of material by a compact object, understanding the nature of a system is very important for studying the evolutionary paths of source populations, and more globally, the evolution of the Galaxy in terms of its source content.\\\\ \\indent \\igr\\ was discovered by \\citet{2009ATel.1997....1T} who reported activity of this source seen with \\integ\\ during the monitoring campaign of \\grs\\ \\citep[e.g.][]{rodrigue08_1915a}. Archival \\swift\\ data of a source named \\object{Swift J1929.8+1818} allowed us to give a refined X-ray position of RA$_{\\mathrm{J}2000}$= 19$^h$ 29$^m$ 55.9$^s$ and Dec$_{\\mathrm{J}2000}$=+18$^\\circ$ 18\\arcmin\\ 39\\arcsec\\ ($\\pm 3.5$\\arcsec) which we used to locate a possible infrared counterpart \\citep{2009ATel.1998....1R}. We also suggested that the \\swift\\ and \\integ\\ sources are the same and that activity from \\igr\\ had been seen with \\swift\\ in the past. The temporal analysis of the XRT light curve showed a possible pulsation at 12.4~s \\citep{2009ATel.1998....1R}. Analysis of \\swift/BAT archival data revealed that the source had been detected on previous occasions with a periodicity at 117.2 days \\citep{2009ATel.2008....1C} that we interpreted as the orbital period of the system. \\\\ \\indent Soon after the discovery of the source with \\integ, we triggered our accepted \\xte\\ and \\swift\\ programmes for follow-up observations of new \\integ\\ sources (PI Rodriguez). A preliminary analysis of the real-time \\xte\\ data allowed us to confirm the pulsations in the signal from the source \\citep{2009ATel.2002....1S} at a barycentred period of 12.44~s indicating that this object hosts an accreting X-ray pulsar. Here we report the detailed analysis of the \\integ, \\swift, and \\xte\\ observations. We begin this paper by detailing the procedures employed for the data reduction. We then describe the results (refined position, X-ray spectral, and temporal analyses) in Sects. 3, 4, 5, and 6, and discuss them in Sect. 7. ", "conclusions": "We have reported here the results obtained from a multi-instrumental campaign dedicated to the X-ray properties of new \\integ\\ sources. The refinement of the X-ray position provided by the \\swift/XRT observations enabled us to identify a possible counterpart at infrared wavelengths. The differences of dereddened magnitudes (Table \\ref{tab:mag}) do not lead to any of the spectral types tabulated in \\citet{tokunaga00}. With \\nh=2.0$\\times 10^{22}$ cm$^{-2}$, J$-$H=0.28 would indicate an F7 V or F8 I star. However the value of H$-$K$_{\\mathrm s}$ is inconsistent with both possibilities. With \\nh=2.73$\\times 10^{22}$ cm$^{-2}$, J$-$H seems too high for any spectral type, although we remark a marginal compatibility (at the edge of the errors on the magnitude) with an O9.5 V star \\citep[J$-$H=$-0.13$, H$-$K$_{\\mathrm s}$=$-$0.04][]{tokunaga00}. This value is, however, a measure of the interstellar absorption through the whole Galaxy. Since \\igr\\ probably lies closer than the other end of the Galaxy, it is likely that the interstellar absorption along the line of sight is lower. In particular, we note that with \\nh=2.5$\\times 10^{22}$ cm$^{-2}$, we obtain J$-$H=$-0.04$ and H$-$K$_{\\mathrm s}$=0.05, which is very close to the values tabulated for a B3 I star \\citep[J$-$H=$-0.03$, H$-$K$_{\\mathrm s}$=0.03][]{tokunaga00}. In that case, the source would lie at a distance $d \\gtrsim 8$ kpc.\\\\ \\indent The X-ray behaviour of the source is indicative of an HMXB: the X-ray spectra are power law like in shape with a hard photon index. One of the spectra (Obs. S1 with the longest exposure) shows evidence for absorption in clear excess to the absorption on the line of sight, which may indicate some intrinsic absorption in this source. However, the evidence is marginal in the other spectra, which could suggest that the intrinsic absorption varies in this system, as has been observed in a number of HMXBs. A long-term periodicity was revealed in the \\swift/BAT data which confirms the binarity of the source \\citep{2009ATel.2008....1C}. We clearly detect an X-ray pulse at a period of 12.44~s, indicating the presence of an X-ray pulsar. We estimate a pulse fraction $P=40\\%$. Apart from millisecond X-ray pulsars, found in systems that are at the end of the evolutionary path of X-ray binaries (and are LMXBs), pulsars are young objects that are usually found in HMXBs. We also remark that the PSD of \\igr\\ shows a broadening at the bottom of the coherent pulsation. Sidelobes and other noise features around coherent pulsation signals have been seen in other HMXBs \\citep[e.g.][in, respectively, \\object{4U 1626$-$67}, and \\object{XTE J0111.2$-$7317}]{kommers98,kaur07}. Such features can be produced by artificial effects (such as the finite length of the time intervals used to make the PSD), or they can be real if, e.g., a quasi-periodic oscillation (QPO) beats with the coherent signal \\citep{kommers98,kaur07}. In the case of \\igr, a possible QPO might be present at $\\sim0.035$~Hz (Fig. \\ref{fig:psd}), even if the quality of the data does not allow us to firmly establish its presence. On the other hand, we cannot exclude that the $0.078$~Hz feature is itself a QPO at a frequency close to that of the coherent pulsation. Given the faintness of the source, we cannot conclude further on that matter. \\\\ \\indent In fact, \\igr\\ lies in a region populated by Be-HMXBs in the so-called ``Corbet diagram'' \\citep{corbet86,2009ATel.2008....1C}, as demonstrated in Fig. \\ref{fig:corbet}. Note that this plot is the most recent update of the Corbet diagram for \\integ\\ sources as of June, 2009 (Bodaghee et al. 2009 in prep.). \\begin{figure} \\epsfig{file=pulse_orbit.ps,width=\\columnwidth} \\caption{Most recent version of the ``Corbet diagram'' including all sources detected with \\integ\\ (Bodaghee et al. 2009 in prep.). Squares represent Be-HMXBs, circles Supergiant-HMXBs (Sg-HMXB), circles in squares are SFXTs, filled symbols represent sources discovered by \\integ. The triangles are HMXBs of unknown nature (including \\igr). The positions of \\igr\\ and of the two SFXTs lying in the Be region of the plot are highlighted. } \\label{fig:corbet} \\end{figure} Recently, \\citet{arash07} further explored the parameter spaces occupied by high-energy sources. They noticed that Be-HMXB and supergiant HMXB also segregate in different parts of the \\nh\\ vs. orbital period and \\nh\\ vs. spin period diagrams. The value of the absorption we obtained through our spectral analysis, combined with the values of the spin and orbital periods, make \\igr\\ lie in a region populated by Be-HMXBs in these diagrams as well. At first order, one can easily understand the segregation in these three different diagrams as results of the age, type of accretion, and probable type of orbit. Be systems are younger, and hence have eccentric orbits, longer orbital periods, and shorter spin periods, whereas supergiant systems are older, are mostly circularised with shorter orbital periods, and longer spin periods. In these latter systems, in addition, the compact object is embedded in the wind of the companion (the feed matter for accretion) which explains their (usually) higher intrinsic absorption. In this respect, all our results point towards \\igr\\ being a Be-HMXB \\citep[the same conclusion, although only based on the Corbet diagram, was presented by][]{2009ATel.2008....1C} .\\\\ \\indent \\integ\\ has unveiled a new (sub-)type of supergiant HMXB, characterised by short ($\\sim$hours) and intense flares seen at X-ray energies: the so-called SFXTs. Models of SFXTs involve stochastic accretion of clumps from the (heterogeneous) wind of the supergiant \\citep[e.g.][]{zand05, negueruela06}, on top of longer, but fainter, periods of activity \\citep[e.g.][]{sidoli07}. In this respect, the detection of periods of short and intense flares in \\igr\\ with \\integ\\ (Fig. \\ref{fig:igrlite}), a behaviour typical of SFXTs, raises the possibility that this source belongs to this new class. Its position in the Corbet and the \\nh\\ vs. spin or orbital period diagrams (Fig.~\\ref{fig:corbet}) is not a definitive contradiction to this hypothesis, since systematic X-ray studies of SFXTs have shown that, contrary to other HMXBs, they seem to populate any part of the diagrams. In particular, \\object{IGR J18483$-$0311} and \\object{IGR J11215$-$5952}, the only two SFXTs for which both orbital and pulse periods are known, lie in the region of Be-systems in these representations (Fig. \\ref{fig:corbet}). The former source has a pulse period of 21.05~s, an orbital period of 18.5~d, and has a B0.5 Ia companion \\citep[see][and references therein]{rahoui08_18483}. The latter has a pulse period of about 187~s, an orbital period of about 165~d \\citep[e.g.][]{sidoli07, ducci09}, and is associated with a B supergiant \\citep{negueruela05}. To reconcile the fact that they are long period systems with a rather low absorption, an eccentric orbit is invoked. \\igr\\ could be the third member of SFXTs (with a known pulse period) having a long and eccentric orbit. In this case, this may point towards the existence of an evolutionary link between Be-HMXBs and eccentric SFXTs (Liu et al. 2009 in prep.). The quality of the data is not good enough to permit us to make any firm conclusions concerning the nature of the system, and only an identification of the optical counterpart's spectral type will resolve this issue." }, "0910/0910.2616_arXiv.txt": { "abstract": "{} {In this paper we study the effects of a toroidal magnetic flux tube emerging into a magnetized corona, with an emphasis on large-scale eruptions. The orientation of the fields is such that the two flux systems are almost antiparallel when they meet.}{We follow the dynamic evolution of the system by solving the 3D MHD equations using a Lagrangian remap scheme.}{Multiple eruptions are found to occur. The physics of the trigger mechanisms are discussed and related to well-known eruption models.} {} ", "introduction": "The trigger mechanism for large-scale eruptions, such as coronal mass ejections (CMEs), is one of the main concerns of theoretical solar physics. There are many different approaches to modelling solar eruptions, ranging from 2D analytical models to 3D numerical simulations. A review of such models can be found in \\citet{forbes00} and further references are given in \\citet{dmac09a}. One highly influential model for solar eruptions is the \\emph{breakout model} \\citep{antiochos99}. This theory uses, as an initial condition, a multipolar configuration, which consists of four distinct flux systems separated by a null point. This configuration is stressed by an \\emph{imposed} shearing at the lower boundary (normally taken to be the photosphere). Reconnection across the current sheet at a stressed null changes the field geometry and weakens the tension of overlying field lines, converting them into nonrestraining field lines of neighbouring flux systems. The continuation of this process can eventually lead to expulsion of a flux rope. For more details of the breakout model, the reader is directed to \\citet{antiochos99}, \\citet{devore08} and \\citet{lynch08}. Recently, \\citet{devore08} and \\citet{lynch08} have simulated magnetic breakout in 3D. The first of these studies finds homologous confined eruptions. The second study follows the topological evolution of a fast breakout CME. \\citet{soenen09} simulate homologous CMEs in the solar wind in an axisymmetric 2.5D configuration. The breakout studies mentioned above have been very useful in investigating the physical mechanisms of solar eruptions. One point which they all have in common, however, is that the initial equilibrium is stressed by artificially imposed shearing motions. Other models are also used to study eruptions in the solar atmosphere. For example, \\citet{mackay06} model the large-scale coronal field as a series of non-linear force force-free equilibria. Flux ropes form but not all of them settle into equilibrium. Those that diverge from equilibrium are ejected. For studying dynamic evolution in the atmosphere, another model is \\emph{flux emergence}. This considers the early evolution of active regions, which, of course, are the sources of large-scale solar eruptions. The standard practice in dynamic flux emergence experiments is to have a stratified solar atmosphere including the top of the solar interior. A flux tube is placed in the solar interior and the system is left to evolve by itself, with no imposed flows. A review of such models can be found in \\citet{archontis08rev}. Some studies include a magnetized corona in their model. \\citet{archontis05} and \\citet{galsgaard07} study the effects of a cylindrical flux tube emerging into a horizontal coronal field. They consider different orientations for the magnetic fields and discuss the reconnection and high-speed jets that occur. \\citet{maclean09} consider the same experiment but investigate the magnetic topology of the system. In relation to solar eruptions, \\citet{manchester04} report on the eruption of a flux rope that forms during the emergence of a cylindrical flux tube into an non-magnetized corona. \\citet{archontis08} study the emergence of a cylindrical flux tube into a magnetized corona and find that with a favourable orientation, a CME-like eruption is possible. In the present paper, we shall consider a similar setup to \\citet{archontis08} but use a different model for the magnetic field. This is a toroidal loop, placed in the solar interior, rather than a cylindrical one \\citep{hood09}. A comparison of these two models is discussed in \\citet{dmac09b}. This paper will show that multiple large-scale eruptions are possible from the same emerging region and will discuss the physics of the trigger mechanisms. The outline of the paper is as follows: $\\S 2$ will describe the model setup and the initial conditions. In $\\S 3$ we shall discuss the results of the eruption experiment and link the processes involved with eruption models. $\\S 4$ will summarize the results. ", "conclusions": "In this paper we have demonstrated that multiple CME-like eruptions are possible from a toroidal flux tube emerging into a magnetized corona. This combines and builds on the work of \\citet{dmac09b} and \\citet{archontis08}. For the present study we consider a corona with a field that is almost antiparallel to the field of the emerging tube. External reconnection at the apex of the emerging arcade weakens the tension of the coronal field. With the expansion of the arcade, this reconnection becomes faster through time. Shearing, which occurs as part of the emergence process, induces internal reconnection in the arcade and produces a flux rope. The continued emergence in combination with removal of the overlying coronal field eventually results in the expulsion of the flux rope. The mechanism for this eruption is similar to that of the breakout model. One important difference, however, is that the external reconnection in this model does not take place at a null point. After the first eruption, continued emergence and, therefore, shearing results in the formation of a second flux rope. This also erupts but the trigger mechanism cannot be directly linked to the breakout model, as with the first eruption. Due to the reconnection of the first eruption with the corona, a weakened coronal field exists above the second rope when it forms. Possible trigger mechanisms, such as runaway reconnection and the torus instability, have been suggested. However, it is possible that the trigger for the second eruption is a combination of such mechanisms." }, "0910/0910.1615_arXiv.txt": { "abstract": "In star formation, magnetic fields act as a cosmic angular momentum extractor that increases mass accretion rates onto protostars and in the process, creates spectacular outflows. However, recently it has been argued that this magnetic brake is so strong that early protostellar disks -- the cradles of planet formation -- cannot form. Our three-dimensional numerical simulations of the early stages of collapse ($\\lesssim 10^5$ yr) of overdense star--forming clouds form early outflows and have magnetically regulated and rotationally dominated disks (inside 10 AU) with high accretion rates, despite the slip of the field through the mostly neutral gas. We find that in three dimensions, magnetic fields suppress gravitationally driven instabilities which would otherwise prevent young, well ordered disks from forming. Our simulations have surprising consequences for the early formation of disks, their density and temperature structure, the mechanism and structure of early outflows, the flash heating of dust grains through ambipolar diffusion, and the origin of planets and binary stars. ", "introduction": "Over the past decade numerical simulations have enabled the exploration of the central physical questions regarding the nature of star formation -- from the collapse of an initial dense gaseous molecular cloud core to the formation of a star and its associated protostellar disk and outflow, and the emergence of a star's planetary system. The formation of disks and jets during star formation is central to many of these issues, but very little is known about the earliest phases of their evolution. How is the initial excessive angular momentum associated with the star's natal core removed -- through magnetic braking \\citep{1994ApJ...432..720B} and then outflows \\citep{2006ApJ...641..949B,2007prpl.conf..277P}, or by spiral density waves in disks \\citep[e.g.~][]{2009arXiv0901.4325L}? Do multiple stars form through gravitational fragmentation of cores or massive disks \\citep{HW2004b,2007MNRAS.377...77P,2008A&A...477...25H}? What is the significance of outflows and jets that are launched before most of the mass has collapsed into the disk \\citep{2003MNRAS.341.1360L, 2006ApJ...641..949B} in a wide variety of young stellar systems -- from brown dwarfs \\citep{2005Natur.435..652W, 2009ApJ...699L.157M} to massive stars \\citep{2007prpl.conf..245A, 2007ApJ...660..479B}? Extraction of angular momentum by magnetic fields that thread the collapsing gas may be too efficient, according to recent two-dimensional \\citep{2008ApJ...681.1356M} and three-dimensional axisymmetric \\citep{2008A&A...477....9H} simulations, preventing the formation of rotationally dominated disks, even when the effects of imperfect coupling of the field with the gas are included \\citep{2009ApJ...698..922M}. These results seemingly contradict the observations which clearly show that disks are present around most if not all young stars, even in environments in which the magnetic field is expected to be strong \\citep{1995AJ....109.1846H}. \\citet{2008A&A...477....9H} performed three-dimensional ideal MHD simulations of collapsing cores using a barotropic equation of state, concluding that no rotationally dominant structure is formed from 10 to 100 AU for highly magnetized cores. \\citet{2008ApJ...681.1356M} performed two-dimensional ideal MHD simulations on collapsing singular isothermal toroids using a barotropic equation of state and an inner boundary at 6.7~AU (effectively a sink particle), finding that even moderately magnetized disks could not form. Such two-dimensional models impose a high degree of mathematical symmetry and therefore miss the formation of bars and spiral waves in disks. We improve on previous results through this three-dimensional adaptive mesh refinement (AMR) investigation which includes a full treatment of the cooling \\citep{2006MNRAS.373.1091B} and the finite coupling of the magnetic field to the pre--stellar gas \\citep[ambipolar diffusion,][]{2008MNRAS.391.1659D}. We find that ordered disk--like structures can emerge on scales $\\lesssim 10$ AU at early times ($\\lesssim 10^5$ yr) in magnetized systems. We have omitted a sink particle in this study as they are expected to affect the solution to within a few sink radii. Without sinks simulations are indeed limited to early disks, although they offer a full solution to the region within 10 AU where the heated core forms. ", "conclusions": "In the early stages of a purely hydrodynamic collapse of a moderately rotating core, material joins the protostar by accreting through a chaotic series of bars and spiral waves. Our results show that in magnetized collapses however, the magnetic field suppresses these wave modes, and a small, regular disk appears at the earliest times. The weakening of magnetic control by ambipolar diffusion is insufficient to guarantee the formation of binary stars in typical cores with moderate rotation. Our results suggest that modest turbulent amplitudes ($>$ 10\\%) appear to be required. Regarding planet formation, the central density structure of early disks (Figure \\ref{fig:2}(e) falls off as $\\Sigma \\propto r^{-(1.7-2.5)}$ much more quickly with disk radius than do protoplanetary disk models at later times, wherein $\\Sigma \\propto r^{-1.5}$ or $r^{-1}$. As the collapse winds up the field early outflows appear -- even for partially coupled disks -- and feed angular momentum and mechanical energy back into the star forming neighborhood. The density and magnetic structure of ideal and partially coupled disks are quite similar -- the main difference is in the strength of the wound up magnetic field (at least an order of magnitude weaker in the ambipolar diffusions case). Finally, we note that evidence for early localized intense drift heating near the disk at the accretion shock, which heats materials up to 1,000K and beyond in a localized region, may be preserved in the observed composition of comets and meteorites. The rapid heating, and subsequent cooling of those crystalline Mg--rich silicate materials that passed through the accretion shock and through the region of high ambipolar heating is reminiscent of heating events that must have occurred for some of these materials seen in cometary grains \\citep{2007prpl.conf..815W}. These events have been attributed to shock heating by spiral waves out to 10 AU in disks \\citep{2005ASPC..341..849D}. As seen in Figure 1(b), only a portion of the accreted disk material will pass through this localized heated region. Subsequent radial turbulent mixing of this flash heated material with the bulk of material that passed through a more gentle heating environment could in principle contribute to the wide mixture of thermal histories preserved in cometary grain materials." }, "0910/0910.3610_arXiv.txt": { "abstract": "We report observations of the remnant of Supernova 1987A with the High Resolution Camera (HRC) onboard the \\emph{Chandra X-ray Observatory}. A direct image from the HRC resolves the annular structure of the X-ray remnant, confirming the morphology previously inferred by deconvolution of lower resolution data from the Advanced CCD Imaging Spectrometer. Detailed spatial modeling shows that the a thin ring plus a thin shell gives statistically the best description of the overall remnant structure, and suggests an outer radius $0.96\\arcsec\\pm0.05\\arcsec \\pm0.03\\arcsec$ for the X-ray--emitting region, with the two uncertainties corresponding to the statistical and systematic errors, respectively. This is very similar to the radius determined by a similar modeling technique for the radio shell at a comparable epoch, in contrast to previous claims that the remnant is 10-15\\% smaller at X-rays than in the radio band. The HRC observations put a flux limit of 0.010\\,cts\\,s$^{-1}$ (99\\% confidence level, 0.08-10\\,keV range) on any compact source at the remnant center. Assuming the same foreground neutral hydrogen column density as towards the remnant, this allows us to rule out an unobscured neutron star with surface temperature $T^\\infty>2.5$\\,MK observed at infinity, a bright pulsar wind nebula or a magnetar. ", "introduction": "The core-collapse supernova \\object{(SN) 1987A} in the Large Magellanic Cloud was the brightest SN observed since the invention of modern telescopes, providing the best opportunity to study the last evolutionary stage of a massive star. Optical observations have revealed a triple-ring nebula centered on the explosion, consisting of an inner equatorial ring of radius 0.81\\arcsec-0.86\\arcsec and two larger outer rings \\citep{bkh+95,plc+95}. These are part of the hourglass-shaped circumstellar medium (CSM) ejected by the progenitor 20,000 years ago \\citep[see][]{mp07}. Since early 2004, the SN blast wave has begun to encounter the main body of inner ring \\citep{pzb+05,pzb+06}. This `big crash' has led to a drastic increase in soft X-ray emission that originates from the optically thin thermal plasma behind the shock. The X-ray spectrum is well-described by a two-component plane-parallel shock model in nonequilibrium ionization, with plasma temperatures $kT\\sim 2$ and 0.3\\,keV, corresponding to the fast and decelerated shocks, respectively \\citep{pzb+04,pzb+06}. As the blast wave propagates into the dense CSM, the soft X-ray emission traces the evolution of the forward shock, probing the CSM structure and its density profile. The nearness of SN~1987A (51.4\\,kpc) allows us to resolve a supernova remnant (SNR) morphology at a very young age. In X-rays, this is only possible with the \\emph{Chandra X-ray Observatory}, the highest resolution X-ray telescope compared to any previous, current and even planned future X-ray missions. Previous \\emph{Chandra} observations of SNR~1987A were all carried out by the Advanced CCD Imaging Spectrometer (ACIS). However, the ACIS detector has a pixel size 0.492\\arcsec, comparable to the full width at half maximum (FWHM) of the mirror's on-axis point-spread function (PSF). Although the effective resolution can be slightly improved by the dithering of the spacecraft and by applying a sub-pixel imaging algorithm \\citep{tmm+01}, image deconvolution is still required to fully resolve the remnant structure \\citep[e.g.][]{bmh+00}. To determine whether artifacts are introduced by the complicated non-linear reconstruction process, the ACIS results need to be compared with higher resolution direct X-ray images. An analogy can be drawn from the radio imaging campaign of SNR~1987A. Since 1992, Australia Telescope Compact Array (ATCA) observations at 9\\,GHz have revealed detailed structure of the radio shell using the super-resolved technique, but it was not until the upgrade of the ATCA in 2003 that the higher resolution diffraction-limited images at 18\\,GHz provided a direct image of the radio morphology \\citep[see review by][]{gsm+07}. The High Resolution Camera (HRC) onboard \\emph{Chandra}, consisting of two microchannel plate detectors, offers smaller electronic readout pixels (0.13175\\arcsec) than ACIS. These better sample the PSF, providing a more straightforward imaging process without the need for deconvolution. Despite a smaller effective area for HRC than ACIS and the lack of any spectral resolution, the SNR is now very bright in X-rays below 2\\,keV, at which the HRC has good sensitivity, making it an ideal instrument for morphological studies. In this Letter, we report a detailed analysis of the first HRC observation of SNR~1987A. \\begin{figure*}[ht] \\centering \\begin{minipage}[c]{0.3\\textwidth} \\includegraphics[height=53mm, bb=5 49 544 503, clip=]{f1a.ps} \\end{minipage} \\hspace*{6.0mm} \\begin{minipage}[c]{0.3\\textwidth} \\includegraphics[height=53mm, bb=89 49 544 503, clip=]{f1b.ps} \\end{minipage} \\hspace*{-3.8mm} \\begin{minipage}[c]{0.3\\textwidth} \\includegraphics[height=53mm, bb=89 49 544 503, clip=]{f1c.ps} \\end{minipage} \\\\[-1pt] \\hspace*{0.1mm} \\begin{minipage}[c]{0.3\\textwidth} \\includegraphics[height=58.7mm, bb=5 1 544 504, clip=]{f1d.ps} \\end{minipage} \\hspace*{5.98mm} \\begin{minipage}[c]{0.3\\textwidth} \\includegraphics[height=58.7mm, bb=89 1 544 504, clip=]{f1e.ps} \\end{minipage} \\hspace*{-3.8mm} \\begin{minipage}[c]{0.3\\textwidth} \\includegraphics[height=58.7mm, bb=89 1 544 504, clip=]{f1f.ps} \\end{minipage} \\caption{(a) Raw HRC data of SNR~1987A taken on 2008 Apr 28-29. The image was binned into the HRC detector pixel and centered at RA=$\\mathrm{05^h35^m28^s}$, Dec=--69\\arcdeg16\\arcmin11.2\\arcsec\\ (J2000). (b) Raw ACIS data taken on 2008 Jan 9. The image is in 0.3-8\\,keV energy range and was binned into the ACIS detector pixel. (c) Super-resolved radio image at 9\\,GHz taken by ATCA on 2008 Apr 23 \\citep{ngs+08}. (d) Deconvolved HRC image using the dataset shown in panel (a). (e) Deconvolved ACIS image using the dataset shown in panel (b), in 0.3-8\\,keV energy range \\citep{rpz+09}. (f) HRC data in panel (a) smoothed to 0.4\\arcsec\\ to match the resolution of the radio image in panel (c). All panels are on the same spatial scale.\\label{f1}} \\end{figure*} ", "conclusions": "We have presented a detailed analysis of \\emph{Chandra} HRC observations of supernova remnant 1987A. The remnant has a similar size as the optical inner ring, and its morphology is statistically best-fitted by a thin ring plus a thin shell. This provides direct evidence that the dense inner ring of circumstellar medium has been overtaken the supernova blast wave. Our spatial modeling also indicates a similar size between the X-ray-- and radio-emitting regions, confirming the picture that the supernova forward and reverse shocks are closely located. The X-ray flux limit on any possible central source rejects a young neutron star, unless it has some fast cooling mechanisms or is obscured by a small accretion disk." }, "0910/0910.1086_arXiv.txt": { "abstract": "{Abundance anomalies observed in globular cluster stars indicate pollution with material processed by hydrogen burning. Two main sources have been suggested: asymptotic giant branch (AGB) stars and massive stars rotating near the break-up limit (spin stars). We propose massive binaries as an alternative source of processed material. We compute the evolution of a 20\\Msun~star in a close binary considering the effects of non conservative mass and angular momentum transfer and of rotation and tidal interaction to demonstrate the principle. We find that this system sheds about 10\\Msun~of material, nearly the entire envelope of the primary star. The ejecta are enriched in He, N, Na, and Al and depleted in C and O, similar to the abundance patterns observed in gobular cluster stars. However, Mg is not significantly depleted in the ejecta of this model. In contrast to the fast, radiatively driven winds of massive stars, this material is typically ejected with low velocity. We expect that it remains inside the potential well of a globular cluster and becomes available for the formation or pollution of a second generation of stars. We estimate that the amount of processed low-velocity material ejected by massive binaries is greater than the contribution of AGB stars and spin stars combined, assuming that the majority of massive stars in a proto-globular cluster interact with a companion and return their envelope to the interstellar medium. If we take the possible contribution of intermediate mass stars in binaries into account and assume that the ejecta are diluted with an equal amount of unprocessed material, we find that this scenario can potentially provide enough material to form a second generation of low-mass stars, which is as numerous as the first generation of low-mass stars, without the need to make commonly adopted assumptions, such as preferential loss of the first generation of stars, external pollution of the cluster, or an anomalous initial mass function. } ", "introduction": "For a long time star clusters have been considered as idealized single-age, chemically homogeneous stellar populations. However, it has recently become clear that many clusters show multiple main sequences and sub giant branches and extended horizontal branches \\citep[e.g.][]{Piotto+07}, implying the existence of multiple populations within one cluster.% In addition, large star-to-star abundance variations are found for light elements such as C, N, O, Na, and Al, while the composition of heavier elements (Fe-group and $\\alpha$-elements) seems to be constant. Field stars with the same metallicity do not exhibit these abundance patterns \\citep[for a review see][]{Gratton+04}. These chemical variations have been interpreted as originating from the presence of both a ``normal'' stellar population, exhibiting abundances similar to field stars of the same metallicity and a second population of stars formed out of material processed by hydrogen burning via the CNO-cycle and by the NeNa and MgAl chains \\citep[e.g.][]{Prantzos+07}. {According to \\citet{Carretta+09}, 50-70\\% of the stars in gloular clusters belong to the second population.} Two sources of processed ejecta have been proposed: the slow winds of \\emph{massive AGB stars}, which enrich their convective envelopes with H-burning products \\citep{Ventura+01,Dantona+02,Denissenkov+Herwig03} and fast-rotating massive stars (we refer to these as \\emph{spin stars}), which are believed to expel processed material centrifugally when they reach break-up rotation \\citep{Prantzos+Charbonnel06,Decressin+07a}.% In this scenario a first generation of stars is formed out of pristine material. Their low-velocity ejecta are trapped inside the potential well of the cluster and provide the material for the formation of a second generation of stars. Although both proposed sources are promising, matching the observed abundance patterns and providing enough ejecta for the formation of a second generation that outnumbers the first generation have proven to be two major challenges. In this Letter we propose \\emph{massive binaries} as a candidate for the internal pollution of globular clusters. \\begin{figure*} \\centering \\includegraphics[angle=-90, width=1.0\\textwidth, bb=50 50 283 616]{13205Fg1.eps} \\caption{ Composition of the slow ejecta of the modeled binary system (Sec.~3) as a function of the ejected amount of mass. The mass fraction $X$ of the main stable isotope of each element is given relative to the initial mass fraction $X_{\\rm i}$, except for Mg where we added $^{24}$Mg$,^{25}$Mg and $^{26}$Mg. The average $X_{\\rm av}$ and the most extreme mass fraction $X_{\\rm ex}$ are given in each panel on a logarithmic scale: $[X] \\equiv \\log_{10} (X / X_{\\rm i})$. For helium we show the absolute mass fraction $Y$ instead. Mass ejected during the first and second mass transfer phases is separated by a thin vertical line. \\label{fig:chem}} \\end{figure*} ", "conclusions": "We propose massive binaries as a source for the internal pollution of globular clusters. The majority of massive stars are expected to be members of interacting binary systems. These return most of the envelope of their primary star to the interstellar medium during non conservative mass transfer. We show that there may be more polluted material ejected by binaries than by the two previously suggested sources: massive AGB stars and the slow winds of fast-rotating massive stars. After dilution with pristine material, as lithium observations suggest, binaries could return enough material to form a chemically enriched second generation that is as numerous as the first generation of low-mass stars, without the need to assume a highly anomalous IMF, external pollution of the cluster or a significant loss of stars from the unenriched first generation. In addition to providing a new source of slowly-ejected enriched material, binary interaction also affects the previously proposed scenarios. Binary mass transfer naturally produces a large number of fast-rotating massive stars that may enrich their surroundings even more. Binary interaction will also affect the yields of intermediate-mass stars. Premature ejection of the envelope in 4-9\\Msun~stars will result in ejecta with less pronounced anti-correlations, as suggested in the AGB scenario. On the other hand, we expect that binary-induced mass loss may also prevent the dredge-up of helium-burning products. For a detailed comparison of the chemical predictions of this scenario, binary models {for a range of masses and orbital periods} are needed and population synthesis models are essential to fullly evaluate the mass budget of the different sources. {Finally, some peculiar feature, such as the apparent presence of distinct, chemically homogeneous subpopulations in $\\omega$~Cen and NGC~2808 \\citep[e.g.][]{Renzini08} deserves further attention. }" }, "0910/0910.0175_arXiv.txt": { "abstract": "The derivation of nebular abundances in galaxies using strong line methods is simple and quick. Various indices have been designed and calibrated for this purpose, and they are widely used. However, abundances derived with such methods may be significantly biased, if the objects under study have different structural properties (hardness of the ionizing radiation field, morphology of the nebulae) than those used to calibrate the methods. Special caution is required when comparing the metallicities of different samples, like, for example, blue compact galaxies and other emission line dwarf galaxies, or samples at different redshifts. ", "introduction": "Why talk about nebular abundances in a symposium devoted to stellar populations in galaxies? The ultimate goal of stellar populations studies is to understand the evolution of galaxies. In principle, the determination of the metallicities of the stellar populations gives an evolutionary view of the chemical enrichment of galaxies. Such studies are just beginning (see e.g. Vale Asari et al. 2009 and contribution to this symposium). For the moment, however, the overwhelming majority of studies on the chemical composition of galaxies rely on global stellar metallicity indices (for early-type galaxies), and on nebular abundances (for star-forming galaxies) derived from emission-line studies. Note that the abundances derived by one or the other method do not have the same meaning: Stellar indices provide an integrated (and weighted) metallicity of the different generations of stars while the chemical composition of nebulae is the result of mixing with the interstellar gas of the metals ejected by \\emph{all} the present and past stellar generations and is the same as that of the presently forming stars. In this communication, we would like to draw attention to widely overlooked biases in the determination of nebular abundances. We will restrict to the so-called ``strong line methods'' for abundance determination (see e.g. Stasi\\'nska 2004 for a general introduction on nebular abundance determinations). After presenting a number of abundance calibrators, We will show how their use may, in certain cases, lead to erroneous statements. ", "conclusions": "Strong line methods to derive nebular abundances are very easy to apply but they are prone to systematic errors. In principle, they should be used only for objects whose HII regions have the same structural proprties as those of the calibrating samples. This recommendation is not easy to follow, but at least one should be aware that using the same calibration for different samples may produces important biases. In particular, claims on differences in oxygen abundance \\begin{itemize} \\item between samples of galaxies with different chemical evolution histories \\item between samples of galaxies with different star formation histories \\item between samples of galaxies at different redshifts (observational selection effects may play a role) \\end{itemize} should be taken with, at a minimum, a grain of salt." }, "0910/0910.0033_arXiv.txt": { "abstract": "Deep \\textit{Swift} UV/Optical Telescope (UVOT) imaging of the Chandra Deep Field South is used to measure galaxy number counts in three near ultraviolet (NUV) filters (uvw2: 1928 \\AA, uvm2: 2246 \\AA, uvw1: 2600 \\AA) and the $u$ band (3645 \\AA). UVOT observations cover the break in the slope of the NUV number counts with greater precision than the number counts by the Hubble Space Telescope (HST) Space Telescope Imaging Spectrograph (STIS) and the \\textit{Galaxy Evolution Explorer} (\\textit{GALEX}), spanning a range from $21 \\lesssim m_{AB} \\lesssim 25$. Number counts models confirm earlier investigations in favoring models with an evolving galaxy luminosity function. ", "introduction": "Galaxy number counts as a function of magnitude provide direct constraints on galaxy evolution in both luminosity and number density. Number counts in the UV, in particular, can help trace the star formation history of the universe. Until recently obtaining faint galaxy number counts in the UV has been difficult due to the small areas surveyed \\markcite{Gardner00,Deharveng94,Iglesias04,Sasseen02,Teplitz06}({Gardner}, {Brown}, \\& {Ferguson} 2000; {Deharveng} {et~al.} 1994; {Iglesias-P{\\'a}ramo} {et~al.} 2004; {Sasseen} {et~al.} 2002; {Teplitz} {et~al.} 2006). While the {\\sl Galaxy Evolution Explorer (GALEX)} has allowed for the measurement of UV galaxy number counts over a wide field of view \\markcite{Xu05}({Xu} {et~al.} 2005, $\\sim 20$ deg$^2$), the confusion limit of {\\sl GALEX} restricts the magnitude range covered to 14 to 23.8 $m_{AB}$. The deepest UV number counts from {\\sl HST\\ } range from $m_{AB}$= 23 to 29 over an extremely small field of view of $\\sim 1.3$ square arcminutes. Here, we present galaxy number counts obtained in 3 near UV filters (1928 \\AA, 2246 \\AA, 2600\\AA) as well as in the $u$ band (3645\\AA) obtained using the {\\sl Swift\\/} UV/Optical Telescope \\markcite{UVOT}(UVOT; {Roming} {et~al.} 2005). Deep exposures were taken of a 289 square arcminute field of view overlapping the Chandra Deep Field South \\markcite{CDFS}(CDF-S; {Giacconi} {et~al.} 2002) allowing for the measurement of number counts from $m_{AB}$ = 21 to 26. UVOT data covers the break in the slope of the NUV number counts with greater precision than the existing {\\sl GALEX} and {\\sl HST} number counts. We use the UVOT number counts to explore the evolution of star-forming galaxies out to $z\\sim1$. ", "conclusions": "" }, "0910/0910.2200_arXiv.txt": { "abstract": " ", "introduction": "Using simple hydrogen atoms, stars forge the entire range of atomic species we observe in Nature. Towards the end of their lives, stars eject these newly forged atoms into interstellar space. Out of this enriched interstellar gas, new stars, their planets and even life form. Stellar mass loss is the motor that makes stars and galaxies change over cosmic time. Yet, when and how mass is lost from stars is still an open debate. Stars more than eight times more massive than the Sun end their lives in spectacular explosions known as supernovae. Stars less than eight times more massive than the Sun lose up to 90\\% of their mass when they become giants, shortly after running of out hydrogen in their cores. The mass that is lost eventually forms beautiful nebulae around the mother star, objects that we call planetary nebulae; a misnomer acquired because they looked like the small circles of planets to 17th century astronomers. Modern telescopes have revealed they exhibit complex shapes (butterfly, multiple lobes emerging from disks, jets and bullets, objects pretty enough to fill coffee table books!). These shapes are in need of an explanation: how do stars lose quite so much mass and why is this mass not residing in a more or less spherical distribution around the star? Simple round shapes have been understood in terms of the Interacting Stellar Winds (ISW) model \\citep{kwok82}. In this model, planetary nebulae (PNe) are formed when stellar material surrounding the central star (CSPN), originating in a slow wind from the asymptotic giant branch (AGB) progenitor, is swept up into a nebular shell by a hot, fast, ionizing post-AGB wind from the white dwarf CSPN. The great majority of PNe are not spherical and many theories have been introduced to explain the wide variety of observed shapes. Many shapes can be reproduced in models by introducing an asymmetry into the slow wind \\citep{kahn85,balick87}. Later theories have concentrated on the origins of the asymmetry, considering stellar rotation and/or magnetic fields \\citep{garcia-segura99,frank04,garcia-segura05}. Most recently though, such effects have been shown to be difficult to maintain in single AGB stars, leaving binary central star systems as the primary progenitors of PNe (see, for example, \\cite{demarco09} for a full review and discussion of the binary progenitor hypothesis). There are several cases though where only the outer shells show a departure from symmetry. This has been postulated to arise from an interaction with the interstellar medium (ISM). In this review, we will consider the current knowledge in this area and the previous work which has led up to the interpretation of the PN--ISM interaction as a four stage evolutionary process, including a period of 'rebrightening', typically occurring late in the lifetime of a PN. We will review this particular part of the process, including the dominating effect of the earlier AGB phase of evolution, and consider the immediate progeny of rebrightened PNe, objects which can bear little resemblance to their origins. The observational support for each stage will also be summarized here for the first time. Finally, we will discuss how rebrightening changes the lifetime estimates of typical PNe and how a sizable sample of interacting PNe may change estimates of the PN population and the understanding of ISM structure on a galactic scale. ", "conclusions": "Simulations have shown that interaction with the ISM strongly affects the death of a PN, even one with an average velocity central star. The structure of the AGB wind around the CSPN is changed from a wall into a bow shock during the preceding phase of evolution. This bow shock and the ram pressure stripping of material into the accompanying tail then dominates the evolution of the PN. Rebrightening and rejuvenation occur, on a short timescale for high speed CSPN, as the PN shell interacts with the bow shock. Following this, the PN is stripped downstream in turbulent regions moving down the tail, resulting in the central star leaving its PN. Further study of rebrightened PN and their surroundings can provide considerable information regarding several other research avenues. As highlighted in the opening paragraph, tracing the ejected mass in circumstellar structures such as bow shocks, tails and late-evolution turbulent regions swept down the tail can reveal much about the way AGB and postAGB stars lose mass. Such information as the position of the bow shock and the parameters of the tail can also reveal, taking into account the method of formation, not only historical stellar mass-loss rates and local ISM conditions, but also details of galactic orbits for the first time. The few objects that have observed lengthy tails, including Mira, HFG1 and Sh 2-68, are currently uniquely placed to derive such orbits. Future surveys of cool dust around AGB stars should reveal many more tails allowing a deeper investigation of a meaningful sample. Once we have a meaningful sample of rebrightened PNe, we should also be able to see how the observable lifetimes of PNe are extended. Current statistical population estimates, which depend on estimated lifetimes, indicate a larger population than we have currently detected, but by extending that lifetime, population estimates accordingly reduce and we are likely to see much more of an agreement between the numbers we actually observe, which is believed to be fairly complete within the Milky Way, and the predicted population. The model of PN--ISM interaction then can account for many characteristics of PNe and explain highly complex, disrupted objects such as Sh 2-68, but it does not reproduce several observed details. In particular, the fragmentation of the nebular rim is not reproduced by the simulations. More advanced models must be able to reproduce this, at least in the cases of high speed CSPN as it seems common at this end of the velocity distribution, e.g. Sh 2-188. No theoretical investigation has yet been performed employing 3D MHD simulations which include the galactic magnetic field. This would seem a natural next step and as Dgani and Soker point out, this may rapidly provide an explanation of particularly fragmentary shells. In a greater challenge to the model, the existence of ancient, effectively spherical PNe with very few signs of ISM interaction are difficult to explain within the model context, especially if they have displaced central stars which rules out an alignment between viewing angle and direction of motion. The model of PN--ISM interaction has come along way and revealed the phenomenon of rebrightened PNe, but there are a still a number of open questions to be solved and no doubt future observations will reveal many more." }, "0910/0910.3206_arXiv.txt": { "abstract": "We present, for the first time, a clear $N$-body realization of the {\\it strong mass segregation} solution for the stellar distribution around a massive black hole. We compare our $N$-body results with those obtained by solving the orbit-averaged Fokker-Planck (FP) equation in energy space. The $N$-body segregation is slightly stronger than in the FP solution, but both confirm the {\\it robustness} of the regime of strong segregation when the number fraction of heavy stars is a (realistically) small fraction of the total population. In view of recent observations revealing a dearth of giant stars in the sub-parsec region of the Milky Way, we show that the time scales associated with cusp re-growth are not longer than $(0.1-0.25) \\times T_{rlx}(r_h)$. These time scales are shorter than a Hubble time for black holes masses $\\mbul \\lesssim 4 \\times 10^6 M_\\odot$ and we conclude that quasi-steady, mass segregated, stellar cusps may be common around MBHs in this mass range. Since EMRI rates scale as $\\mbul^{-\\alpha}$, with $\\alpha \\in [\\frac{1}{4},1]$, a good fraction of these events should originate from strongly segregated stellar cusps. ", "introduction": "The distribution of stars around a massive black hole (henceforth MBH) is a classical problem in stellar dynamics \\citep{1976ApJ...209..214B,1977ApJ...211..244L}. The observational demonstration of the existence of nuclear stellar clusters (henceforth NSCs)---as revealed by a clear upturn in central surface brigthness---in the centers of galaxies makes its study ever more timely. A number of NSCs in coexistence with a central MBH have recently been detected \\citep{2009MNRAS.397.2148G} suggesting that NSCs around MBHs, like the one in the center of the Milky Way, may be quite common. The renewed interest in this theoretical problem is thus motivated by the observational data in NSCs and, in particular, the very rich and detailed data available for the stars orbiting the Galactic MBH. At the same time, the prospects for detection of gravitational waves (GWs) from extreme mass ratio inspirals (henceforth EMRIs) with future GW detectors such as the {\\it Laser Interferometer Space Antenna} (LISA) also urge us to build a solid theoretical understanding of sub-parsec structure of galactic nuclei. In fact, EMRI rates will depend strongly on the stellar density of compact remnants as well as on the detailed physics within $O$($0.01$pc) of the hole, which is the region from which these inspiralling sources are expected to originate \\citep{2005ApJ...629..362H}. \\cite{1976ApJ...209..214B} have shown, through a kinetic treatment that, in the case all stars are of the same mass, this quasi-steady distribution takes the form of power laws, $\\rho(r) \\sim r^{-\\gamma}$, in physical space and $f(E) \\sim E^p$ in energy space ($\\gamma=7/4$ and $p=\\gamma-3/2=1/4$). This is the so-called {\\it zero-flow solution} for which the net flux of stars in energy space is precisely zero. \\cite{2004ApJ...613L.109P} and \\cite{2004ApJ...613.1133B} were the first to report $N$-body realizations of this solution, thereby validating the assumptions inherent to the Fokker-Planck (FP) approximation---namely, that scattering is dominated by uncorrelated, $2$-body encounters and, in particular, dense stellar cusps populated with stars of the {\\it same mass} are robust against ejection of stars from the cusp. The properties of stellar systems that display a range of stellar masses are only very poorly reproduced by single mass models. It is well known from stellar dynamical theory that when several masses are present there is mass segregation---a process by which the heavy stars accumulate near the center while the lighter ones float outward \\citep{1987degc.book.....S}. Accordingly, stars with different mass get distributed with different density profiles. By assuming a stellar population with two mass components, \\cite{1977ApJ...216..883B}---hencefort BW77---generalized their early cusp solution and argued heuristically for a scaling relation $p_L = m_L/m_H \\times p_H$ that depends on the star's mass ratio only. However, they obtained no general result on the inner slope of the heavy objects; nor did they discuss the dependence of the result on the component's number fractions. On the other hand, it was shown long ago by \\cite{1969A26A.....2..151H} that the presence of a mass spectrum leads to an increased rate of stellar ejections from the core of a globular cluster, but he did not include the presence of a MBH at the center. H\\'enon's work raises the question as to whether {\\it multi-mass} stellar cusps, obtained from the solution of the FP equation, are robust against ejection of stars from the cusp. Ejections---due to strong encounters---are {\\it a priori} excluded from the FP evolution, even though they could occur in a real nucleus. Furthermore, even if cusps were shown by $N$-body results to be robust against stellar ejections (and we show that they are in this Letter), BW77 scaling cannot be valid for arbitrary number fractions. Recently \\cite{2009ApJ...697.1861A}---henceforth AH09---stressed this latter point and have shown via FP calculations that, indeed, in the limit where the number fraction of heavy stars is realistically small, a new solution that they coined {\\it strong mass segregation} obtains with density scaling as $\\rho_H(r) \\sim r^{-\\alpha}$, where $\\alpha \\gtrsim 2$. They have shown that there are two branches for the solution parametrized by $\\Delta = \\frac{N_HM_H^2}{N_LM_L^2} \\cdot \\frac{4}{3+M_H/M_L}$. The $weak$ branch, for $\\Delta > 1$ corresponds to the scaling relations found by BW77; while the $strong$ branch, for $\\Delta < 1$, generalizes the BW77 solution. There is a straightforward physical interpretation. In the limit where heavy stars are very scarce, they barely interact with each other and instead sink to the center due to dynamical friction against the sea of light stars. Therefore, a quasi-steady state forms in which the heavy star's current is not nearly zero and thus the BW77 solution does not hold. As $\\Delta$ increases, self-scattering of heavies becomes important and the resulting quasi-steady state forms with a nearly zero current for stars of all masses, so BW77 solution is recovered. For all these reasons, it is fundamental to verify the Bahcall-Wolf solution---as well as its Alexander-Hopman generalization---with $N$-body integrations. There has been a surprisingly small number of $N$-body studies of multi-mass systems around a MBH \\citep{2004ApJ...613.1143B,2006ApJ...649...91F}, and none of them reported the occurence of strong mass segregation. In this Letter we use direct $N$-body integrations to show for the first time that: (i) strong mass segregation is a robust outcome of the growth of stellar cusps around a MBH when $\\Delta < 1$; (ii) BW77 solution is recovered when $\\Delta > 1$; (iii) as a corollary, we conclude that the rate of stellar ejections from the cusp is too low to destroy the high density cusps around MBHs---even though ejections from the cusp {\\it do} occur. Furthermore, having validated the FP formalism, we proceed to use it to estimate the time scales for cusp re-growth starting from a wider range of models. \\cite{2009arXiv0909.1318M} obtained, for Milky Way type nucleus, times in large excess of a Hubble time. We use a FP formalism which, in contrast with that of the latter author, is tailored to follow the simultaneous evolution of the cusp of different stellar masses without any restrictions with respect to the values of $f(E)$ or $\\rho(r)$. With our FP solutions we show that, for $\\mbul \\lesssim 5 \\times 10^6 M_\\odot$, the times for re-growing stellar cusps are shorter than a Hubble time. Our results clearly suggest that strongly segregated stellar cusps around MBHs in this mass range may be quite common in NSCs and should be taken into account when estimating EMRI event rates. ", "conclusions": "Our results show that strong mass segregation is a {\\it robust} outcome from the growth of stellar cusps around MBHs. We have used $N$-body integrations with two masses---light and heavy components representing main sequence stars and stellar BHs respectively---, and compared the results with those obtained with the FP formalism. The broad agreement between both methods validates the FP description of the {\\it bulk} properties of time-evolving stellar distribution around a MBH---and its underlying assumptions. The differences of quantitative detail are the subject of another work in preparation. Using the FP equation to study cusp growth under a variety of initial conditions purported to represent cored nuclei, we have shown that the time scales associated with cusp re-growth are clearly shorter than a Hubble time for nuclei with MBHs in the mass range $\\mbul \\lesssim 5 \\times 10^6 M_\\odot$---even though the relaxation time, as estimated for a single mass stellar distribution, exceeds a Hubble time in the upper part of this mass range. Therefore, our work strongly suggests that quasi-steady---strongly segregated--- stellar cusps may be common around MBHs with masses in this range. EMRIs of compact remnants will be detectable by LISA precisely for MBHs in this mass range \\citep{2006PhRvD..74b3001D,2007CQGra..24..113A,2007PhRvD..75b4005B}. Estimates for event and detection rates by LISA costumarily assume that the stellar cusps are in steady state \\citep{2006ApJ...645.1152H,2006ApJ...645L.133H}. But recent observations reveal a dearth of giants inside $1$ pc from $SgrA^*$ and raise the possibility that cored nuclei are common---this scenario has been thoroughly explored by \\cite{2009arXiv0909.1318M}. Our results strongly suggest that stellar cusps can re-grow in less than a Hubble time. The existence of cored nuclei still remains plausible though---especially for nuclei with MBHs in the upper part of the mass range---, since time scales are still quite long ({\\it e.g.} $6$ Gyr in Milky Way type nuclei). However, since EMRI rates scale as $\\mbul^{-\\alpha}$, $\\alpha \\in [\\frac{1}{4},1]$, and re-growth times are $\\lesssim 1$ Gyr for $\\mbul \\lesssim 1.2 \\times 10^6 M_\\odot$, we still expect that a substantial fraction of EMRI events will originate from segregated stellar cusps. Finally, indirect observations alone will reveal whether there is a ``hidden'' cusp of old stars and their dark remnants around $SgrA^*$ \\citep{2005ApJ...622..878W,2009ApJ...703.1743P}. \\ \\ \\ \\ \\" }, "0910/0910.3947_arXiv.txt": { "abstract": "X-ray observations of several active galactic nuclei show prominent iron K-shell fluorescence lines that are sculpted due to special and general relativistic effects. These observations are important because they probe the space-time geometry close to distant black holes. However, the intrinsic distribution of Fe line strengths in the cosmos has never been determined. This uncertainty has contributed to the controversy surrounding the relativistic interpretation of the emission feature. Now, by making use of the latest multi-wavelength data, we show theoretical predictions of the cosmic density of relativistic Fe lines as a function of their equivalent width and line flux. We are able to show unequivocally that the most common relativistic iron lines in the universe will be produced by neutral iron fluorescence in Seyfert galaxies and have equivalent widths $< 100$~eV. Thus, the handful of very intense lines that have been discovered are just the bright end of a distribution of line strengths. In addition to validating the current observations, the predicted distributions can be used for planning future surveys of relativistic Fe lines. Finally, the predicted sky density of equivalent widths indicate that the X-ray source in AGNs can not, on average, lie on the axis of the black hole. ", "introduction": "\\label{sect:intro} The X-ray spectrum of many active galactic nuclei (AGNs) exhibits a strong emission feature at an energy of 6.4~\\kev\\ that is due to the fluorescence of weakly ionized iron in a dense and relatively cold medium \\citep[e.g.,][]{nan89,pou90,np94}. The strength of the emission feature requires that the region responsible for the fluorescence subtend about half the sky as seen from the illuminating X-ray source \\citep[e.g.,][]{gf91}. As the variability properties of AGN place the X-ray source within 10~Schwarzschild radii from the black hole \\citep[e.g.,][]{gra92,utt07}, it is likely that relativistic effects may alter the observed line shape \\citep{fab89,laor91} allowing it to be a powerful probe of space-time curvature. Several examples of relativistically broadened \\fe\\ lines have been detected in the spectra of AGNs over the last decade \\citep[e.g.,][]{tan95,fab02,rn03,jmm07,fab09}. These detections are leading to measurements of black hole spins and accretion disk dynamics \\citep{br06,rf08}. The strength of the Fe line emission depends on only four parameters: (i) the shape of the illuminating continuum \\citep{bfr02,br02}, which in AGNs is a power-law with photon index $\\Gamma$; (ii) the abundance of iron (relative to the solar value) in the accretion disk \\citep{bfr02}, $A_{\\mathrm{Fe}}$; (iii) the ionization parameter of the illuminated disk \\citep{bfr02}, $\\xi=4\\pi F_{X}/n_{\\mathrm{H}}$, where $F_{X}$ is the illuminating flux, and $n_{\\mathrm{H}}$ is the density of the illuminated slab; and (iv) the reflection fraction $R$, a measure of the relative strength of the reflection component in the observed spectrum (this is related to the covering factor of the accretion disk). Recently, multi-wavelength survey data has been able to measure correlations between each of the first three of these parameters to $\\lambda=L_{\\mathrm{bol}}/L_{\\mathrm{Edd}}$, the Eddington ratio of AGNs \\citep{rye09,ith07,nt07}. Here, $L_{\\mathrm{bol}}$ is the bolometric luminosity of an AGN, and $L_{\\mathrm{Edd}} = 1.3\\times 10^{38}\\ (M_{\\mathrm{BH}}/M_{\\odot})$~erg~s$^{-1}$ is the Eddington luminosity of a black hole with mass $M_{\\mathrm{BH}}$ and $M_{\\odot}$ is the mass of the Sun. In addition, various surveys have been able to measure the black hole mass distribution \\citep{net09} and its evolution with redshift \\citep{lab09}, as well as the luminosity function of AGNs as a function of redshift \\citep[e.g.,][]{ueda03}. We can then combine all this information to calculate the density and evolution of the \\fe\\ line from accretion disks. In the following we concentrate solely on the \\fe\\ line that arises from the inner accretion disk, and neglect the contribution from any narrow component \\citep[e.g.,][]{nan06}, or one that arises from reflection off a Compton-thick absorber \\citep[e.g.][]{my09}. The following $\\Lambda$-dominated cosmology is assumed \\citep{sper03}: $H_0=70$~km~s$^{-1}$~Mpc$^{-1}$, $\\Omega_{\\Lambda}=0.7$, and $\\Omega_{m}=0.3$. ", "conclusions": "\\label{concl} This work has shown, for the first time, the cosmic density and EW distribution of relativistic \\fe\\ lines. Sensitive observations of these line profiles are vital as they allow measurements of fundamental parameters such as the spin of the central black hole. Our results validate the previous and current results from \\textit{XMM-Newton} and \\textit{Suzaku} that have shown that the majority of relativistic \\fe\\ lines have EWs$< 100$~eV \\citep{nan07}. Now that models of the intrinsic line EW and flux distributions are available, detailed planning of future \\fe\\ surveys by \\textit{IXO} can be performed. The sensitivity of \\textit{IXO} will push spin measurements beyond the local universe and out to moderate redshifts, revolutionizing our understanding of the cosmic evolution of black holes. Finally, we have shown that the extreme light-bending model predicts many more intense relativistic \\fe\\ lines than the $R=1$ model (Figs. 1 and 2), in disagreement with current observational constraints. This implies that, on average, the X-ray source in AGNs does not lie on the axis of the black hole where light-bending would be so extreme. It seems more likely that the X-ray source in average AGNs lies above the accretion disk at a radius of several $r_g$ from the black hole where light bending is less severe. Comparing data from future surveys of \\fe\\ lines with plots such as Figure~3 should allow a measurement of the average radial displacement of the X-ray source. Of course, this conclusion does not preclude the possibility that extreme light-bending occurs in a few rare sources." }, "0910/0910.3481_arXiv.txt": { "abstract": "We present a fully consistent evolutionary disc model of the solar cylinder. The model is based on a sequence of stellar sub-populations described by the star formation history (SFR) and the dynamical heating law (given by the age-velocity dispersion relation AVR). The stellar sub-populations are in dynamical equilibrium and the gravitational potential is calculated self-consistently including the influence of the dark matter halo and the gas component. The combination of kinematic data from Hipparcos and the finite lifetimes of main sequence (MS) stars enables us to determine the detailed vertical disc structure independent of individual stellar ages and only weakly dependent on the IMF. The disc parameters are determined by applying a sophisticated best fit algorithm to the MS star velocity distribution functions in magnitude bins. We find that the AVR is well constrained by the local kinematics, whereas for the SFR the allowed range is larger. The model is consistent with the local kinematics of main sequence stars and fulfils the known constraints on scale heights, surface densities and mass ratios. A simple chemical enrichment model is included in order to fit the local metallicity distribution of G dwarfs. In our favoured model A the power law index of the AVR is 0.375 with a minimum and maximum velocity dispersion of 5.1\\,km/s and 25.0\\,km/s, respectively. The SFR shows a maximum 10 \\,Gyr ago and declines by a factor of four to the present day value of 1.5\\,$\\msun/\\mathrm{pc}^2/\\mathrm{Gyr}$. A best fit of the IMF leads to power-law indices of -1.46 below and -4.16 above 1.72$\\msun$ avoiding a kink at 1$\\msun$. An isothermal thick disc component with local density of $\\sim 6\\%$ of the stellar density is included. A thick disc containing more than 10\\% of local stellar mass is inconsistent with the local kinematics of K and M dwarfs. Neglecting the thick disc component results in slight variations of the thin disc properties, but has a negligible influence on the AVR and the normalised SFR. The model allows detailed predictions of the density, age, metallicity and velocity distribution functions of MS stars as a function of height above the mid-plane. The complexity of the model does not allow to rule out other star formation scenarios using the local data alone. The incorporation of multi-band star count and kinematic data of larger samples in the near future will improve the determination of the disc structure and evolution significantly. ", "introduction": "\\label{introd} There is no 'concordance' Milky Way model available so far that describes the structure, kinematics, chemistry and evolution of the disc in great detail. The classical stellar density model of \\citet{bah80a,bah80b,bah84c} composed by a spheroid and an exponential disc with magnitude-dependent exponential scale heights is still widely used. \\citet{bah84a,bah84b} introduced a finite set of isothermal disc components and solved the Poisson and the Jeans equation for a dynamical equilibrium model of the vertical disc structure. At present the most sophisticated Milky Way model based on star counts is the so-called 'Besan\\c{c}on model' of the Galaxy developed and presented in a series of articles. In \\citet{rob03} a general description and the present status of the model is given. In this model the thin disc is composed by a sequence of stellar sub-populations with increasing age and velocity dispersion. The Besan\\c{c}on model has still some serious drawbacks in the construction of the thin disc concerning the density profiles of the components, the star formation history (SFR) and the initial mass function (IMF) that will be discussed below. The main input functions to be determined in an evolutionary disc model are the SFR and the dynamical heating described by the age-velocity dispersion relation (AVR). The $\\mathrm{SFR}(t)$ of the Milky Way disc is still not very well determined. The main reason for that is the lack of good age estimators with corresponding unbiased stellar samples. Especially samples selected by colour cuts or by a magnitude limit are biased with respect to the age distribution, because there is an age-metallicity relation due to the chemical enrichment of the disc. Therefore not even the famous Geneva-Copenhagen sample of 14,000 F and G stars \\citep{nor04,hol09} (hereafter GCS1,GCS2) with well determined stellar properties and individual age estimates can be used to derive the star formation history directly by star counts. Systematic biases introduced by the method of GCS1 concerning age and stellar parameter determinations are discussed in \\citet{pon05} and \\citet{hay06}. The Besan\\c{c}on model of the Galaxy was developed and presented in a series of articles. In \\citet{hay97a,hay97b} the SFR of the thin disc is determined to be approximately constant using a series of isothermal components and applying an approximate method to achieve dynamical equilibrium. In further applications they used a constant SFR combined with a steep IMF in the sensitive mass regime of 1-3\\,$\\msun$ \\citep{rob03}. As a result, the mean age of the disc population, the scale heights and the surface densities of the stellar sub-populations are relatively small. \\citet{roc00} used chromospheric age determinations of late-type dwarfs. They apply stellar evolution and scale height corrections. The main result is the determination of enhanced star formation episodes over the lifetime of the disc. They exclude a constant SFR from chemical evolution models. In \\citet{fue04} the star formation history for open star clusters was determined. They found at least five episodes of enhanced star formation in the last 2\\,Gyr. Since star clusters contain only a small percentage of all disc stars, an extrapolation to the total (smoothed) SFR is not possible. In recent years the Hipparcos stars with precise parallaxes and proper motions were used to determine the SFR with different methods. \\citet{her00} determined the SFR over the last 3\\,Gyr using isochrone ages. They found a series of star formation episodes on top of an underlying smooth SFR. This result is complementary to our model, which gives the slow evolution of the smoothed SFR. \\citet{ver02} applied an inverse method to derive from the colour magnitude diagram the star formation history and age-metallicity relation. They used prescribed AVR and gravitational potential and fixed the IMF to be a power law over the whole stellar mass range. This ansatz results in a steep IMF and a rather young thin disc. Additionally they found strong maxima in the SFR at ages of $10^7$ and $2\\times10^9$\\,yr. In \\citet{bin00} and in \\citet{cig06} the scale height variation of main sequence stars was not taken into account. Therefore these models derive the local age distribution of stellar sub-populations with lifetimes larger than the age of the disc instead of the SFR. \\citet{bin00} determined the age of the solar neighbourhood to be $11.2\\pm 0.75$\\,Gyr, which includes the contribution of the thick disc. Both papers found an approximately constant local age distribution, which corresponds to a decreasing SFR according to the increasing scale height with increasing age of the stellar sub-populations. In a recent paper \\citet{aum09} derived the SFR using the updated Hipparcos data \\citep{vle07} and the kinematics of the GCS sample. They calculated in detail the number of main sequence (MS) stars as function of B-V colour based on a KTG93-like IMF and scale height corrections due to dynamical heating. Their favoured disc model has an exponentially decaying SFR with a decay timescale of 8.5\\,Gyr and an age of 12.5\\,Gyr. In the model the gravitational potential is fixed and the SFR depends strongly on the properties of the IMF by construction. The presented paper (Paper I) starts a new attempt to develop a fully consistent disc model which is able to predict the properties of the Milky Way disc in great detail concerning density distributions, age distributions, kinematics and chemistry of all types of stars. The aim of this sequence of papers is to construct a smooth physically consistent disc model that can be extended to a global Galaxy model and allows detailed predictions of number densities and kinematics of stars of different types. The new model can be of some help for the construction of a 'concordance' model of the Milky Way. It can be used for the preparation of the future astrometric space mission Gaia that will measure with high precision positions, distances, proper motions and stellar parameters (temperature, surface gravity and chemical composition) of one billion stars \\citep{per01,bai05}. We start with a local disc model of the solar cylinder, which is based on the SFR, the AVR and a chemical evolution function described by the age-metallicity-relation (AMR). The vertical density profiles and the corresponding scale height determinations of thin and thick disc stars are calculated in a self-consistent gravitational potential including the DM halo and the gas component. The model parameters are determined by a best fit procedure of the velocity distribution functions $f_\\mathrm{i}(|W|)$ of solar neighbourhood stars from B--K dwarfs. We use Hipparcos stars and at the faint end the Catalogue of Nearby Stars (CNS4), select the MS stars and divide these into a series of volume complete sub-samples in magnitude bins. The derived $f_\\mathrm{i}(|W|)$ of each sub-sample are compared simultaneously to the model predictions. Details of the fitting procedure and the significance with respect to parameter variations are discussed in a future paper \\citep{jus09} (Paper II). Since the velocity distribution function of MS samples is the average over the lifetime of these stars properly weighted by the SFR and the dilution due to the increasing scale height, the time resolution is strongly limited. Therefore we use only smooth input functions for the SFR, AVR and metal enrichment in the model. The resulting SFR, AVR and chemical evolution describe the long-term smoothed disc evolution. Similar disc models by fitting the vertical luminosity and colour profiles of edge-on galaxies instead of the kinematics have already been used successfully \\citep{jus96,jus06}. The main advantages of this method is that it does not depend on individual stellar age determinations and that it is essentially independent of the shape of the IMF. A weakness is that we cannot confine the SFR strongly based on local data only. The application of the model to star counts of remote stellar populations will be of great help in this respect. In subsequent work the local model will be continuously extended and compared to large data samples taken from the Sloan Digital Sky Survey SDSS/SEGUE \\citep{aba09} and the Radial Velocity Experiment RAVE \\citep{zwi08} to further constrain the possible parameter range of the disc model. A major restriction of a local model is that radial mixing of stars in the disc cannot be included easily. There are two main mixing processes discussed in the literature. The first one is directly connected to the dynamical heating process responsible for the AVR. The gravitational scattering process leads to a diffusion of orbits in velocity space (AVR) and in position (radial, tangential, vertical mixing). \\citet{wie96} discussed the radial diffusion of stellar orbits and the consequences for the AMR in some detail. The radial probability distribution function of the birth-places depends mainly on the age of the stellar sub-population and only weakly on the physical scattering process. For a 5\\,Gyr old population the typical radial width of the initial distribution is $\\pm 2$\\,kpc. The second mechanism is resonant scattering of circular orbits by spiral arms \\citep{sel02,ros08,sch09}. This mechanism changes the radial position of stars quickly, but the eccentricity of the orbits remain very small. The effect of resonant scattering may account for up to half the stars in the solar neighborhood \\citep{ros08}, which shows the possible scale of the errors which may result if migration is ignored. Unfortunately the efficiency and the properties of resonant scattering in the Milky Way disc are still poorly known. In a future global disc model radial mixing may be included in a parametrized form. The local model presented in this paper is primarily a description of the 'status quo' of the local Milky Way disc quantified by the local age distribution (SFR), the kinematics (AVR) and the chemical composition (AMR). Therefore the relations between these function are not altered by radial mixing. But the physical interpretation of the SFR as a 'local star formation history' and the AVR as a 'local dynamical heating process' are weakened by radial mixing and require corrections. For the chemical evolution model the consequences are more severe. The tight connection of the enrichment, the SFR and the gas infall will be broken. The chemical enrichment of the gas decouples partly from the (apparent) enrichment of the stellar population. A more general local AMR including an intrinsic scatter would be the consequence as derived by \\citet{sch09}. In section \\ref{model} the physics of the disc model is presented, section \\ref{observ} describes the observational data, in section \\ref{disc} the fitting procedure and the properties of the best fit models are given, section \\ref{summary} collects the main results and discussion. ", "conclusions": "\\label{summary} We presented a new disc model for the thin disc in the solar cylinder based on a continuous star formation history (SFR) and a continuous dynamical heating law (AVR) of the stellar sub-populations. This new model combines and improves the advantages of several different attempts to model the vertical structure in the solar neighbourhood: We used a sequence of isothermal sub-populations in dynamical equilibrium as \\citet{bah84c} and \\citet{aum09} did; We used the full velocity distribution functions $f|W|$ as \\citet{hol00} did, and not only the velocity dispersions (the AVR); We solved the Poisson equation self-consistently including the thick disc, gas and DM halo contribution as done in the Besan\\c{c}on model \\citep{rob03}. A chemical evolution model with reasonable gas infall rate, which was tuned to reproduce the local [Fe/H] distribution of G dwarfs, is included. This enables us to apply correct MS luminosities and lifetimes. Additionally our model is insensitive to the IMF, because it is based on the normalised velocity distribution functions of MS stars. We determined pairs of (SFR, AVR) by a best fit of the local kinematics. The SFRs which are consistent with the MS velocity distribution functions show a decline factor below five down to unity (= const. SFR). The strongest feature that would distinguish between a constant and a declining SFR is a direct determination of the age distribution of low mass stars in the solar neighbourhood. Despite the large variety of SFRs there is a strong correlation to the AVR. For each SFR the slope and maximum velocity dispersion of the AVR are well determined. For the AVR we find a power law with indices between 0.375 and 0.5. The range of models is consistent with the results of \\citet{bin00} for the local age distribution and \\citet{aum09} for the SFR. Applying the stellar lifetimes and the new scale height corrections to the PDMF results in an IMF that shows only one break point at $1.7\\msun$ and a steep falloff at high masses. The density profile of an isothermal thick disc component can be fitted very good by a sech$^\\alpha_\\mathrm{t}$ profile, where $\\alpha_\\mathrm{t}$ depends on the velocity dispersion. The most prominent effect of thick disc stars is the enhancement of the high velocity wings of K and M dwarfs in the range of 40--60\\,km/s. From that we can exclude a heavy thick disc with distinct kinematics and more than 10\\% contribution to the local stellar density. Changing the thick disc parameters leads to slight variations of the thin disc properties (mainly by assigning part of the stellar disc to the thick disc), but has a negligible influence on the normalised SFR and AVR of the thin disc. A variety of predictions can be made from the new disc model. The density profiles of MS star sub-populations depend on the lifetime of the stars and are significantly different to density profiles of single age sub-populations. The shape is neither exponential nor sech$^2$ and can be characterised by the (half-)thickness and the exponential scale height. From the vertical density profiles MS stars number densities as function of colour and apparent magnitude can be predicted. Applying these number densities with observed 'Hess' diagrams from large surveys like the catalogues of the Sloan Digital Sky Survey SEGUE/SDSS enables us to restrict the parameters of the SFR further. We determined vertical gradients in the kinematics which will be tested with Radial Velocity Experiment (RAVE) data. Age and metallicity distributions of stellar sub-populations as a function of $z$ above the galactic plane are predicted. The future plan is to extend the local model to a complete disc model of the Milky Way that provides a fully self-consistent connection of stellar densities and kinematics. Ultimately this kind of detailed model is essential to understand the large data sets as already available from SDSS and which are expected on a much higher level in amount and precision by PanSTARRS and the astrometric Gaia satellite mission." }, "0910/0910.3162_arXiv.txt": { "abstract": "We have derived nebular abundances for 10 dwarf galaxies belonging to the M81 Group, including several galaxies which do not have abundances previously reported in the literature. For each galaxy, multiple H~\\ii\\ regions were observed with GMOS-N at the Gemini Observatory in order to determine abundances of several elements (oxygen, nitrogen, sulfur, neon, and argon). For seven galaxies, at least one H~\\ii\\ region had a detection of the temperature sensitive [OIII] $\\lambda$4363 line, allowing a ``direct\" determination of the oxygen abundance. No abundance gradients were detected in the targeted galaxies and the observed oxygen abundances are typically in agreement with the well known metallicity-luminosity relation. However, three candidate ``tidal dwarf'' galaxies lie well off this relation, UGC~5336, Garland, and KDG~61. The nature of these systems suggests that UGC 5336 and Garland are indeed recently formed systems, whereas KDG~61 is most likely a dwarf spheroidal galaxy which lies along the same line of sight as the M81 tidal debris field. We propose that these H~\\ii~regions formed from previously enriched gas which was stripped from nearby massive galaxies (e.g., NGC~3077 and M81) during a recent tidal interaction. ", "introduction": "Due to their low dust content, simple kinematics, modest to negligible abundance gradients, and prevalence in the universe, dwarf galaxies are excellent objects to use as probes of galaxy evolution (e.g., Hunter \\& Gallagher 1985; Kobulnicky \\& Skillman 1997). Additionally, these small galaxies may be remnants of the building blocks that combined to form larger galaxies early in the history of the universe (e.g., Bullock \\& Johnston 2005). Thus, understanding the chemical evolution of dwarf galaxies is critical for constraining galaxy evolution models. Indeed, numerous studies have investigated dwarf galaxy abundances (e.g., Lequeux et al.\\ 1979; Skillman et al.\\ 1989; Richer \\& McCall 1995; Kunth \\& \\\"Ostlin 2000 and references therein). These studies have established the well known metallicity-luminosity relation, wherein more massive galaxies have higher oxygen abundances. While this empirical trend has been measured repeatedly, its underlying cause(s) remains controversial. On the one hand, greater retention of enriched gas by the deeper potential wells of more massive galaxies could produce the relation (e.g., Gibson \\& Matteucci 1997). However, another possibility is that pristine gas is processed more efficiently by larger galaxies. Thus, a combination of gas transport (i.e., inflows and outflows) and systematic variations in star formation efficiency could produce this trend (e.g., Dalcanton 2007). At the same time, most galaxies do not exist in isolation. Evolution and gas transport are influenced via galaxy-galaxy interactions; such influence is likely quite extreme in some cases. If galaxies are hierarchically built structures, such interactions may be the {\\sl de facto} method of galaxy evolution. Thus, interacting galaxies may further inform us as to how the composition of the universe is evolving. Has interaction significantly altered the chemical composition of systems? The nearby M81 Group (D $\\sim$ 4~Mpc) offers an ideal laboratory in which to investigate the signatures of chemical enrichment in a dynamic environment. This prominent group is one of the nearest galaxy associations to the Local Group. Hence, even low mass group members may be observed and included in studies. Additionally, the proximity of the M81 Group has permitted accurate distance determinations (Karachentsev et al.\\ 2002 and references therein) and resolved studies of the stellar populations (e.g., Weisz et al.\\ 2008). As recently as 300 Myrs ago, the primary galaxies of the M81 Group experienced a dramatic collision (e.g., van der Hulst 1979; Yun et al.\\ 1994). The tidal HI debris still connects the three most massive galaxies involved in the event: M81, M82, and NGC 3077. There have been detailed studies of this tidal debris in both HI and CO (Taylor et al. 2001, Chynoweth et al. 2008, Brinks et al. 2008). Additionally, several clumps of new star formation, likely induced by the three-body interaction, have been identified (e.g., Davidge 2008). Indeed, it has been suggested that some dwarf galaxies within the group have recently been formed through tidal interactions (Makarova 2002; Sabbi et al. 2008; Weisz et al.\\ 2008). Accordingly, one may hypothesize that the metallicity of such tidally formed galaxies could be elevated, compared to other galaxies of similar mass (e.g., Weilbacher et al. 2003); the gas from which they formed may be pre-enriched by the larger system from which the gas was stripped. To investigate this possibilty, we have obtained spectra for 10 of the dwarf galaxies of the M81 Group, including some of the ``tidal dwarf\" candidates, to determine the intrinsic metallicity of these systems. Several of these galaxies do not have previously determined abundances in the literature. We present these data and discuss their implications for the evolution of the M81 Group. In \\S2 we present the observations and data processing. The emission line measurements and abundance determinations are discussed in \\S3 and \\S4, respectively. The results of the dwarf irregular galaxies are discussed in the context of the M81 Group in \\S5, while tidal dwarf galaxies are discussed in \\S6. Lastly, \\S7 summarizes our findings. We adopt 12$+$log(O/H) = 8.93 (Anders \\& Grevesse 1989) as the solar value of the oxygen abundance for the present discussion. \\section {Observations} \\subsection{Optical Spectroscopy} Low-resolution spectra of ten M81 Group dwarfs were obtained\\footnote[1]{Program ID GN-2006A-Q-26.} in queue mode with the Gemini Multiple Object Spectrograph (GMOS) instrument (Hook et al. 2004) on Gemini North during May 2006. Galaxies were selected to span a large range of both star formation rate and optical luminosity. Global parameters for the selected targets are listed in Table~\\ref{t:param}. Masks were cut to place 5\\arcsec\\ long slitlets with a 1.5\\arcsec\\ slit-width on regions observed to emit significant flux in the H$\\alpha$ pre-imaging observations. Blue spectra were obtained using the B600 (600 lines mm$^{-1}$) diffraction grating with a resolution of 0.45 $\\rm{\\AA}$ pixel$^{-1}$. Red spectra were obtained using the R600 diffraction grating with a resolution of 0.47 $\\rm{\\AA}$ pixel$^{-1}$. Spectral observations were typically 2$\\times$900s for both blue and red setups. We used $4 \\times 2$ spectral-spatial binning, respectively, resulting in a spectral resolution of 1.8 $\\rm{\\AA}$ per binned pixel. Observations were centered at 4500 $\\rm{\\AA}$ and 4550 $\\rm{\\AA}$ in the blue, and 7000 $\\rm{\\AA}$ and 7050 $\\rm{\\AA}$ in the red, to ensure full spectral coverage across the two detector gaps. We did not use an order-blocking filter for any of the spectra. This yielded complete spectral coverage from roughly 3500 $\\rm{\\AA}$ to 8000 $\\rm{\\AA}$ for the majority of slitlets. The astrometric slit positions of emission sources presented in this work are listed in Table~\\ref{t:positions}. Five apertures revealed regions of strong hydrogen emission yet displayed no sign of oxygen lines; these sources were not analyzed as part of this work as they do not seem to be H~\\ii\\ regions, but their positions are also listed in Table~\\ref{t:positions}. The position angle of the slitlets and the average parallactic angle for each pointing are also given in Table~\\ref{t:positions}. Due to scheduling constraints, it was not always possible to line up the slits with the parallactic angle to reduce the light loss from atmospheric refraction. The fields most affected by this misalignment are UGC 4459, UGC 5139, UGC 5918, and UGC 8201. The spectra were reduced and analyzed using the IRAF\\footnote[2]{IRAF is distributed by the National Optical Astronomical Obsevatories.} GMOS package. Reduction of spectra included bias subtraction and flat fielding based on observations of the quartz halogen lamp on the GCAL unit. A spectral trace of a bright continuum source was used to define the trace for all slits in each field. The shape of this trace was consistent in all exposures of a given field. Wavelength calibration was obtained by observations of CuAr arc lamps interspersed between observations of the target galaxies. As all of the galaxies in this sample have relatively large angular extents, most filled a large fraction of the $5\\arcmin \\, \\times \\, 5\\arcmin$ GMOS-N field of view. Thus, the edges of each 5\\arcsec\\ slit lie well within the glow of the diffuse ionized medium of the host galaxy. Furthermore, some H {\\sc ii} regions are quite extended and completely fill these short 5\\arcsec\\ slitlets. Therefore, we investigated whether standard local sky subtraction procedures could be used for these observations, given the concern that slits include both sky and diffuse ionized gas. Indeed, local sky subtraction often over-subtracted strong lines relative to weaker emission lines, as expected since the diffuse ionized gas is at a different temperature than the compact H~\\ii\\ region. Even in the most compact H {\\sc ii} regions observed, a 10\\% reduction in the flux of strong lines was found when the sky was measured locally. Thus, a more accurate sky background was measured in slits located away from the main body of the target galaxy. Average red and blue sky spectra were created by combining multiple one dimensional spectra from slits sampling the sky for every galaxy observed. While the same physical region was extracted from both the blue and red setups, the size of this extraction region varied from slit to slit. Thus, the amount of sky background in each slit was different, as it depended on the spatial extent of the slit with significant line emission. To ensure the proper level of sky background was subtracted, spectral regions populated by sky lines and free of any substantial extraterrestrial emission line were used as reference regions (e.g., 7150--7300 \\AA). Average sky spectra were scaled such that residuals in the reference regions were minimized when object and sky spectra were differenced. Subsequently, the scaled backgrounds were subtracted from the one-dimensional H{\\sc ii} region spectra. One-dimensional spectra were corrected for atmospheric extinction and flux calibration based on observations of the flux standard Feige 67 (Oke 1990). The flux standard was observed on each night science data were obtained. A representative flux-calibrated spectrum is shown for each target galaxy in Figure \\ref{fig:spec}. Note that red and blue continuum levels are in excellent agreement, despite observations being non-simultaneous. This confirms stable sky conditions and indicates that the extraction regions were well matched in both the red and blue setups. These example spectra are of similar quality to other spectra used for the abundance determinations of each target. \\subsection{Line intensities} The strength of emission features were measured in the one-dimensional spectra and subsequently corrected for Balmer absorption and line of sight reddening. Reddening estimates were determined by comparing the ratios of flux from Balmer emission lines. Ratios of intrinsic line strength were interpolated from the tables of Hummer \\& Storey (1987) for case B recombination, assuming T$_e$=12500 K and N$_e$=100 cm$^{-3}$. When the temperature sensitive [O {\\sc iii}] $\\lambda$4363 line was detected, temperatures derived from the [O {\\sc iii}] lines were used rather than 12500 K. The Galactic reddening law of Seaton (1979) parameterized by Howarth (1983) was applied to derive the reddening function normalized at H$\\beta$, assuming R=A$_V$/E$_{B-V}$=3.1. When the measured reddening coefficient, c$_{H\\beta}$, differed for the H$\\alpha$/H$\\beta$ and H$\\gamma$/H$\\beta$ line ratios, an underlying Balmer absorption of up to 2${\\rm \\AA}$ was applied. The determined values of c$_{H\\beta}$ are reported in Table~\\ref{t:lineratiob}. Reddening corrected line intensities measured from H {\\sc ii} regions in the target fields are reported in Tables~\\ref{t:lineratiob} and \\ref{t:lineratior}. The regions are grouped according to the field with which they are associated, as listed in the first column. The error associated with each measurement is determined from measurements of the Poisson noise in the line measurement, error associated with the sensitivity function, Poisson noise in the continuum, read noise, sky noise, flat fielding calibration error, error in continuum placement, and error in the determination of the reddening. Note that while the spectral coverage permitted measurements of lines redward of the [Ar {\\sc iii}] $\\lambda$7136 line, the numerous skylines made measurements in the red portion of the spectra very uncertain. Furthermore, both [O {\\sc i}] lines in our spectral region are coincident with strong atmospheric emission features. Occasionally, the resulting profile of the sky-corrected [O {\\sc i}] $\\lambda$6300 and $\\lambda$6364 lines had broad wings indicative of over subtraction. Thus, measurements of [O {\\sc i}] lines are not reported when sky subtraction did not yield trustworthy line profiles. \\subsection{Diagnostic Diagrams} Since the red and blue spectra were not obtained simultaneously and, further, queue scheduling did not always permit observations along the parallactic angle, diagnostic diagrams were examined to verify we recovered reliable line ratios, and hence reliable abundances. For example, the [N~\\ii]/H$\\alpha$ and [O~\\iii]/H$\\beta$ sequence is shown in Figure \\ref{fig:diag}a along with a theoretical curve from models of H~\\ii\\ regions (Baldwin, Phillips, \\& Terlevich 1981). H~\\ii~regions from the current study are shown as circles and pentagons, while the triangles indicate H~\\ii~regions from a study of spiral galaxies (van Zee et al.\\ 1998). Following McCall et al.\\ (1985), Figure \\ref{fig:diag}b shows the same diagnostic, replacing [O~\\iii]/H$\\beta$ with the sum of [O~\\ii]/H$\\beta$ and [O~\\iii]/H$\\beta$, commonly referred to as R$_{23}$. Our data are in good agreement with both theoretical predictions and the general locus of galaxies in all diagnostic diagrams. Intriguingly, H~\\ii~regions from the observed candidate tidal dwarfs, KDG 61, Garland, and UGC 5336, do not occupy the same parameter space as H~\\ii~regions from the dwarf irregulars in our sample. Rather, both diagnostic trends indicate they resemble H~\\ii~regions observed in larger spiral galaxies. We discuss these targets in more detail in \\S5. \\section {Nebular Abundances} Determination of elemental abundances from spectra of ionized gas requires determination of (i) the electron density (n$_e$), (ii) the electron temperature (T$_e$), and (iii) an estimate of the abundance of atomic species in unobserved ionic states. In spectra where the [O {\\sc iii}] $\\lambda$4363 line is detected, we may directly follow the methodology of Osterbrock \\& Ferland (2006), since this line is highly sensitive to the electron temperature. We adopt the standard practice of referring to this as the ``direct\" method (Dinerstein 1990). When measurement of the weak line is impossible, we have employed a strong line method wherein an oxygen abundance is deduced from photoionization models and measurements of the strong [O {\\sc ii}] $\\lambda$3727 and [O {\\sc iii}] $\\lambda$5007,4969 lines (e.g., Edmunds \\& Pagel 1984, McGaugh 1991). Subsequently, a consistent electron temperature is deduced based on the strong-line oxygen abundance (van Zee et al.\\ 1997). While the absolute measurement of this ``semi-empirical\" approach is less certain, it still provides robust relative abundances since relative abundances are less dependent on the electron temperature. \\subsection{Direct Abundance Determinations} Osterbrock \\& Ferland (2006) carefully describe the procedure to determine the electron density and the electron temperature of an H~\\ii\\ region. We mention here only the most relevant points in the determination of T$_e$ and n$_e$. The [S {\\sc ii}] $\\lambda$6717/6731 line ratio is sensitive to the electron density of the H {\\sc ii} region. The majority of H {\\sc ii} regions studied here are within the low-density limit of this diagnostic ratio [$I(\\lambda6717)/I\\lambda6731)>1.35$]. Accordingly, an electron density of 100 cm$^{-3}$ was assumed for most H {\\sc ii} regions. In the six cases where the line ratio did not lie below the low-density limit, electron densities were calculated from the observed [S {\\sc ii}] ratio using a version of the FIVEL program (De Robertis 1987). Detection of the [O~\\iii] $\\lambda$4363 line in a spectrum permits a direct determination of the electron temperature of the ionized gas. We used the [O~\\iii] $\\lambda$4363 line for a temperature measurement only when the S/N in the line exceeded 2.8. Following this threshold for detection, we were able to measure reliable reddening corrected [O~\\iii] $\\lambda$4363 fluxes for 26 H~\\ii~regions (56\\% of the sample). These 26 H~\\ii\\ regions were distributed among 7 of the 10 observed fields. From these fluxes, electron temperatures were determined for the O$^{++}$ region of the nebula. Subsequently, the electron temperature in the O$^+$ region was calculated from the O$^{++}$ electron temperature using the approximation of Pagel et al.\\ (1992). With electron densities and temperatures in hand, emissivity coefficients for detected emission lines were determined using the FIVEL program (De Robertis et al. 1987). While emission lines were detected for all dominant ionization states of oxygen, derivation of abundances for other elements requires accounting for the presence of atomic species whose ionization states were unobserved. Nitrogen abundances were derived under the assumption N/O=N$^+$/O$^+$; neon abundances were derived under the assumption Ne/O = Ne$^{++}$/O$^{++}$ (Peimbert \\& Costero 1969). In the cases of both sulfur and argon, the analytical ionization correction factors of Thuan, Izotov, \\& Lipovetsky (1995) have been adopted. When the [S~\\iii] $\\lambda$6312 line was not detected, the sulfur abundance was deemed to be too uncertain to report. \\subsection{Semi-empirical Abundance Determinations} In the 20 H~\\ii\\ regions where the [O~\\iii] $\\lambda$4363 line could not be cleanly measured or was simply too weak to be detected, we are unable to determine the electron temperature via a direct approach. However, the strong oxygen lines ([O~\\ii] $\\lambda$3729 and [O~\\iii] $\\lambda$4959,$\\lambda$5007) are sensitive to both oxygen abundance and electron temperature. Therefore, a semi-empirical approach was used, where oxygen abundances were estimated from the strong line ratio (e.g., McGaugh 1991); we uniformly adopt an error of 0.20 for abundances determined via the semi-empirical method as discussed in van Zee et al.\\ (2006). In this approach, where photoionization models are used as calibration, the geometry of the H~\\ii\\ region is represented by the average ionization parameter, \\=U, the ratio of ionizing photon density to particle density, as determined by the ratio of [O~\\iii] to [O~\\ii]. This additional parameter increases the spread of abundance estimates for a given R$_{23}$. The theoretical photoionization models of McGaugh (1991) are shown in Figure \\ref{fig:mcg} along with the observed line ratios presented in this work; solid points indicate the [O~\\iii] $\\lambda$4363 line was detected while open points denote spectra where the line was unmeasurable. In general, our data lie in the region predicted by theoretical models; however, three points fall off the model grid. Two of these regions, KDG 61-9 and Garland-5, have detections of the [O {\\sc iii}] $\\lambda$4363 line; therefore, the electron temperature was determined directly from line ratios and does not rely upon photoionization models. The remaining point, UGC 5336-11, is from a supernova remnant. As such, it would not be expected to follow predictions of pure photoionization but must be compared to models of shock ionization (see the discussion on UGC 5336 in \\S4). The strong-line ratio of ([O~\\ii]+[O~\\iii])/H$\\beta$, or $R_{23}$, is a smooth function that depends on the electron temperature and oxygen abundance of a gas; however, $R_{23}$ is double valued. Collisionally excited Lyman series emission accounts for the cooling of gas in a very low metallicity H~\\ii\\ region. As the metal content increases, contributions to cooling from infrared fine structure lines become more important. Around an oxygen abundance of one-third of the solar value (12$+$log(O/H) $\\sim$ 8.4), cooling becomes dominated by fine structure lines, causing the R$_{23}$ surface to fold over despite the increased metallicity. The fold in the $R_{23}$ surface is shown in Figure \\ref{fig:mcg}. The fold-degeneracy may be broken using the [N~\\ii] to [O~\\ii] line ratio as a diagnostic (e.g., Alloin et al.\\ 1979). In the majority of cases here, H~\\ii\\ regions exhibited log([N~\\ii]/[O~\\ii])$<-1.0$. Accordingly, the ``lower-branch'' and thus lower values for the oxygen abundance were adopted for those H~\\ii\\ regions. The high abundance value was generally adopted for regions in the ``upper-branch'' with log([N~\\ii]/[O~\\ii])$>-1.0$. However, some ($\\sim5$) of the observed H~\\ii\\ regions fell near this critical value. In such instances, the two possible strong-line abundances were compared to abundances from other H~\\ii\\ regions within the field. As the interstellar-medium chemical abundances of dwarf galaxies are generally spatially homogeneous (e.g., Kobulnicky \\& Skillman 1997; Lee \\& Skillman 2004; Lee et al.\\ 2005, 2006), the abundance branch yielding a more consistent value with other spectra was adopted; often these other spectra had detections of $\\lambda$4363, eliminating this uncertainty on their behalf. The selected H~\\ii\\ regions in UGC 5692 were all faint targets with large errors in the [N~\\ii] to [O~\\ii] ratio. Thus, this diagnostic line ratio did not present a clear resolution to the R$_{23}$ degeneracy. While no spectra from UGC 5692 yielded a solid [O~\\iii] $\\lambda$4363 detection, two of the apertures had weak detections of this line. These detections were more consistent with higher temperature H~\\ii\\ regions. Furthermore, the temperatures derived from the [O~\\iii] line ratios yielded oxygen abundances consistent with the lower branch of the R$_{23}$ relation. Therefore, the lower abundance was adopted for H~\\ii\\ regions in UGC 5692. The strong-line abundance calibrations we have used are derived from zero-age H~\\ii\\ regions (McGaugh 1991). As an H~\\ii\\ region ages, the characteristic spectrum and ionization parameter may evolve, introducing systematic abundance errors (e.g., Stasi\\'nska \\& Leitherer 1996; Olofsson 1997). van Zee et al.\\ (2006) showed that such corrections may be significant when log([O~\\iii]/[O~\\ii])$<-0.4$ in an H~\\ii\\ region. Within the present sample, two regions in Garland (Garland-6 and Garland-9) and three regions in UGC 5336 (UGC5336-3, UGC5336-11, and UGC5336-12) lie within this highly uncertain parameter space. Weisz et al. (2008) have shown concentrations of blue stars in both of these systems, indicative of star formation in the last 100 Myr. Thus, these HII regions are associated with recent star formation, but we have no constraint on the character of the radiation field (i.e., reflective of a fully populated zero-age main sequence versus an evolved population of lower mass main sequence stars). Lacking more information on the H~\\ii\\ regions sampled by Garland-6, Garland-9, and UGC5336-3, we have made no corrections for the age of the nebulae in the current analysis. Abundance determinations from UGC5336-11 and UGC5336-12 are discussed further in \\S4. \\subsection{Abundances} The abundance determinations for each H {\\sc ii} region are presented in Table~\\ref{t:abund}. Abundances were calculated using the measured line strengths and corresponding emissivity coefficients as determined by FIVEL (De Robertis et al.\\ 1987). Errors associated with the derived abundances were clearly dominated by the uncertainty of the electron temperature. The abundances of nitrogen and the $\\alpha$-elements (Ne, S, and Ar) relative to oxygen are shown in Figure \\ref{fig:alpha}, along with solar values from Anders \\& Gervasse (1989). At higher oxygen abundance, H~\\ii\\ regions do show an increased N/O ratio. This relative increase is consistent with secondary production of nitrogen dominating the primary component at higher metallicities (see Figure 4 of Vila-Costas \\& Edmunds 1993). As expected for primary elements, $\\alpha$/O trends appear to be constant as oxygen abundance increases. The mean log (Ne/O) is $-0.82\\pm0.12$; the mean log (Ar/O) is $-2.21\\pm0.13$; the mean log (S/O) is $-1.50\\pm0.10$. These mean values are in agreement with large samples of H {\\sc ii} regions in dwarf galaxies (e.g., Thuan et al.\\ 1995; Izotov et al.\\ 2004; Lee et al.\\ 2004; van Zee et al.\\ 2006). By placing multiple apertures on targeted galaxies, we are able to investigate the uniformity of oxygen abundance across dwarf galaxies in the M81 Group. Except for the candidate tidal dwarfs, the standard deviation of oxygen abundances within each galaxy is smaller than the error associated with the individual measurements. This reaffirms previous results which have shown no detectable abundance gradients within dwarf galaxies (e.g., Kobulnicky \\& Skillman 1996, 1997). While the suspected tidal dwarf galaxies Garland and UGC 5336 may show real spatial trends in oxygen enrichment, these galaxies are not representative of undisturbed dwarf galaxies. These dwarf galaxies and their origins are discussed further in \\S6.1~--~\\S6.3. The average oxygen abundance and log(N/O) for a given galaxy are tabulated in Table~\\ref{t:param}. The average abundance was determined with a weighted average of the individual measurements. This relies upon the assumption that each H~\\ii\\ region is an independent measure of the composition of the galaxy, which is taken to be uniform. Abundances determined using the direct method are more certain. Thus, when direct and semiempirical determinations were available, the strong line abundances were not used in the average. It should be noted that all strong line abundances agree, within the errors, with the average abundances determined for their respective galaxies. Also, due to their larger uncertainties, the semi-empirical abundance determinations would have very little effect in weighted averages. \\section {Comparison with Previous Work} Previous studies of the M81 Group have reported abundances for half of the galaxies in this sample. We compare our findings with those of previous studies below. In general, we note that the new results presented here are consistent with previous observations of these systems. {\\it UGC 4459.} -- Hunter \\& Gallagher (1985), Skillman, Kennicutt \\& Hodge (1989), Hunter \\& Hoffman (1999), Pustilnik et al.\\ (2003) and Saviane et al.\\ (2008) all obtained spectra of UGC 4459. They determined an oxygen abundance, 12+log(O/H), of 8.59, 7.62, 7.79, 7.52$\\pm$0.08 and 7.83, respectively, where [O~\\iii] $\\lambda$4363 was only detected by Pustilnik et al.\\ (2003). While the values of Skillman, Kennicutt \\& Hodge (1989) and Pustilnik et al.\\ (2003) are lower than the value determined here using the direct method (7.82$\\pm0.09$), they are in excellent agreement with the semi-empirical abundance we calculate for UGC 4459 (7.61$\\pm0.20$). We adopt the mean oxygen abundance determined from the direct method in this paper, which is in excellent agreement with both the Hunter \\& Hoffman (1999) and Saviane et al.\\ (2008) results. {\\it UGC 5139.} -- A strong line abundance is reported for UGC 5139 based only on the detection of $\\lambda$4959 and $\\lambda$5007 (Miller \\& Hodge 1996). To derive an abundance, they estimated the strength of [O~\\ii] based upon theoretical ratios of [O~\\iii]/[O~\\ii]. Our semi-empirical result for this same H~\\ii\\ region (7.75$\\pm0.20$) is in excellent agreement with their value of 12+log(O/H)$= 7.7\\pm0.3$. Here, we adopt our direct-method average abundance of 8.00$\\pm0.10$. {\\it UGC 4305.} -- Oxygen abundances of 12+log(O/H)=8.55, 7.92, and 7.71$\\pm$0.13 were reported by Hunter \\& Gallagher (1985), Masegosa et al.\\ (1991), and Lee et al.\\ (2003), respectively. While the strong line abundance of Hunter \\& Gallagher (1985) appears to have been placed on the ``upper-branch\" of the strong line abundance models, the three HII regions where the [OIII] $\\lambda$4363 line was detected by Masegosa et al.\\ (1991) are in excellent agreement with our derived value of 7.92$\\pm$0.10. We note the presence of strong He \\ii\\ $\\lambda$4686 emission in UGC 4305-11, which may indicate the presence of another ionizing source, such as a Wolf-Rayet star. Such a composite spectrum could feature blends of lines which would not follow the general properties of a photoionized region. Thus, we do not include UGC 4305-11 in our average abundance of UGC 4305. {\\it UGC 5666.} -- Multiple H~\\ii\\ regions in UGC 5666 were studied by Masegosa et al.\\ (1991) and Miller \\& Hodge (1996), yielding 12 + log(O/H) = 8.09 and 8.06, respectively; these results are from a combination of direct and empirical methods. While the current results are very tightly clumped together, with a resultant average of 7.93 $\\pm$ 0.05, they lie $\\sim0.1$~dex below the previously reported abundances. UGC 5666 extends well beyond the $5\\arcmin \\, \\times \\, 5\\arcmin$ field of view of GMOS-N. Therefore, diffuse galactic emission in the sky slit could have produced a systematic offset as discussed in \\S2.1. However, it is also worth noting Masegosa et al.\\ (1991) obtained their data in 1984, before linear CCDs were widely employed, and thus a discrepancy of only 0.1 dex may not be significant. {\\it UGC 5336.} -- An investigation of the optical counterpart of an X-ray source led to the acquisition of spectra from a U-shaped object north-east of UGC 5336 (Miller 1995). Two of our apertures located in the UGC 5336 field lie upon this object. While Miller (1995) reported anomalous H$\\alpha$/H$\\beta$ ratios for these regions and thus did not correct for line-of-sight reddening, our new observations yield normal H$\\alpha$/H$\\beta$ ratios indicative of foreground reddening. In agreement with Miller (1995), we find the [S~\\ii] $\\lambda$6717,6731/H$\\alpha$ for slits UGC 5336-10 and UGC 5336-11 are 0.6 and 0.7 respectively, clearly above the criterion necessary for an object to be classified as a supernova remnant (Skillman 1985; Smith et al.\\ 1993). Since UGC 5336-10 and UGC 5336-11 are not from purely photoionized H~\\ii\\ regions, proper abundance determinations require shock models. Thus, diagnostic plots using [N~\\ii ]/H$\\alpha$, [S~\\ii] $\\lambda$6731/H$\\alpha$ and [O~\\iii]/H$\\beta$ have been constructed using the MAPPINGS III radiative shock ionization models (Allen et al.\\ 2008). Under the assumption of a simple low density shock model, both diagnostics indicate 12+log(O/H)$\\sim8.4$. Given that this value is consistent with the oxygen abundance results of photoionization models, we adopt the semi-empirical abundance determinations for these apertures. The abundance determined from apertures sampling the supernova remnant (8.44$\\pm0.20$) as well as UGC 5336 (8.86$\\pm0.20$) are in excess of the findings of Miller (1995: 12+log(O/H)$\\sim$8.0). While these two objects may not be physically associated, they appear to be linked by a common H I envelope. Accordingly, we include measurements from all four apertures in the abundance determinations for UGC 5336. ", "conclusions": "" }, "0910/0910.3212_arXiv.txt": { "abstract": "\\noindent We present a parametric analysis of the intracluster medium and gravitating mass distribution of a statistical sample of 20 galaxy clusters using the phenomenological cluster model of \\citeauthor{ascasibar08}. We describe an effective scheme for the estimation of errors on model parameters and derived quantities using bootstrap resampling. We find that the model provides a good description of the data in all cases and we quantify the mean fractional intrinsic scatter about the best-fit density and temperature profiles, finding this to have median values across the sample of 2 and 5 per cent, respectively. In addition, we demonstrate good agreement between \\rfiveh\\ determined directly from the model and that estimated from a core-excluded global spectrum. We compare cool core and non-cool core clusters in terms of the logarithmic slopes of their gas density and temperature profiles and the distribution of model parameters and conclude that the two categories are clearly separable. In particular, we confirm the effectiveness of the logarithmic gradient of the gas density profile measured at 0.04\\rfiveh\\ in differentiating between the two types of cluster. ", "introduction": " ", "conclusions": "" }, "0910/0910.5328_arXiv.txt": { "abstract": "{The \\agile{} \\gray satellite accumulated data over two years on several blazars. Moreover, for all of the sources detected by AGILE, we exploited multi wavelength observations involving both space and ground based telescopes and consortia, obtaining in several cases broad-band spectral energy distributions (SEDs) which span from the radio wavelengths up to the TeV energy band. I will review both published and yet unpublished \\agile{} results on \\gray blazars, discussing their time variability, their \\gray flare durations and the theoretical modeling of the SEDs. I will also highlight the GASP-WEBT and \\swi{} fundamental contributions to the simultaneous and long-term studies of \\gray blazars. ", "introduction": "Multi wavelength studies of \\gray active galactic nuclei (AGNs) date back to the late '70s and the early '80s with the COS--B detection of 3C~273 \\citep{Swanenburg78, Bignami81}. Nevertheless, the paucity of extragalactic \\gray source detected by SAS-2 and COS-B prevented systematic multi frequency studies. It was during the '90s, with the launch of \\cgro{}, that \\egret{} allowed to establish blazars as a class of \\gray emitters and to start multi wavelength studies of such sources. For a few sources, it was possible to study both the properties of the SEDs during different \\gray states, and the search for correlated variability at different bands, as for 3C~279 \\citep{Sed_Hartman2001, Var_Hartman2001}. The recent launches of the \\agile{} and \\fermi{} satellites allowed the blazar community to observe a large fraction of the sky above 100~MeV, thanks to their wide ($\\sim 3$\\,sr) field of view (FoV), and to start a more effective multi wavelength approach in their spectral energy distribution investigation. In the following, I will briefly introduce the \\agile{} satellite, and then I will focus on the \\agile{} results on the studies of \\gray blazars. Particular emphasis will be given to the importance of simultaneous (or at least, co-ordinated) multi wavelength observations, in order to study both the broad-band properties, and the correlations between the emission at different frequencies. ", "conclusions": "During its first year of sky monitoring, \\agile{} demonstrated the importance of its wide ($\\sim 3$\\,sr) field of view in detecting transient sources at high off-axis angles. Moreover, its unique combination of a \\gray detector with a hard X-ray monitor allowed us to study in detail the high energy portion of the blazar SEDs. The synergy between the \\agile{} wide field of view, its fast response to external triggers, and the availability of a network of ground-based telescopes, allowed us to obtain a multi wavelength coverage for almost all the detected sources, and to investigate the physics of different classes of blazars. Moreover, by means of long-term studies of selected objects, we were able to monitor both high and low \\gray states of different sources. Finally, archival data analysis is in progress, and we start detecting dim and steady sources \\citep{Pittori2009AA:catal}." }, "0910/0910.0482_arXiv.txt": { "abstract": "We describe a simple test of the spatial uniformity of an ensemble of discrete events. Given an estimate for the point source luminosity function and an instrumental point spread function (PSF), a robust upper bound on the fractional point source contribution to a diffuse signal can be found. We verify with Monte Carlo tests that the statistic has advantages over the two-point correlation function for this purpose, and derive analytic estimates of the statistic's mean and variance as a function of the point source contribution. As a case study, we apply this statistic to recent gamma-ray data from the \\Fermi Large Area Telescope (LAT), and demonstrate that at energies above 10 GeV, the contribution of unresolved point sources to the diffuse emission is small in the region relevant for study of the WMAP Haze. ", "introduction": "Statistical tests of isotropy have a long history in astronomy. A common question is ``What fraction of the observed emission could originate from unresolved point sources?'' For example, possible point source contributions to the extragalactic X-ray background were investigated by \\cite{1974MNRAS.166..329S}, and the small-angle power spectrum of the cosmic far infra-red background has been used to estimate the isotropic component \\citep{1996ApJ...473L...9K}. More recently the Auger team tested the isotropy of ultra-high energy cosmic ray events by cross correlating with positions of known active galactic nuclei (AGN; \\cite{Cronin:2007zz,Abraham:2007si}) to provide information on their origin. However in the absence of an appropriate external catalog, such cross-correlation methods cannot be used, motivating consideration of a more general approach. In some cases a detector provides binned counts (e.g. pixels in a CCD); in other cases, photon event directions are reconstructed in some other way (e.g. a gamma-ray pair conversion telescope). In the latter case, it is desirable to apply statistics that do not require binning of the data, as binning introduces additional arbitrary parameters into the problem. In the limit of low flux density, where the mean density of photon events (hereafter, ``events'') is much less than one per PSF, explicit detection of point sources may become impractical, and estimation of the unresolved point source flux becomes especially difficult. In some cases, the two-point correlation function, or some modified form \\citep[e.g.][]{Ave:2009id}, is used as a test. However, the Fourier transform of a field of point sources has significant phase correlation, and a two-point function (or a power spectrum) discards this phase information. Higher order correlation statistics capture it, but are somewhat cumbersome to use. In the following, we describe a statistic that is easy to understand and evaluate, and is optimized to address this question, particularly in the case of fairly sparse data sets with (on average) $\\lesssim 1$ event per PSF circle. The key insight is that if a substantial fraction of the photons come from point sources, it is much more likely that two photons appear within one PSF of each other than in the diffuse case. This is true even if the expected integrated flux is of order one count. Likewise, the number of PSF circles containing \\emph{no} counts is larger if point sources contribute. These considerations motivate us to define a ratio between the fraction of ``isolated events'' and the fraction of ``empty circles.'' This ratio is very closely related to the fraction of diffuse emission, and can be calibrated with Monte Carlo simulations for specific choices of instrumental parameters and a putative luminosity function. The two-point function, in contrast, is weighted by the density squared and is \\emph{not} proportional to the desired quantity. In the following sections we define the statistic, estimate its variance, show how it behaves in various limits, generalize it to the case where many events appear in every PSF circle, and show a practical application to recently released data from the \\Fermi Gamma-ray Space Telescope. \\begin{figure*} \\includegraphics[width=0.48\\textwidth]{plots/fsratio-diff100.ps} \\includegraphics[width=0.48\\textwidth]{plots/fsratio-diff85.ps} \\includegraphics[width=0.48\\textwidth]{plots/fsratio-diff50.ps} \\includegraphics[width=0.48\\textwidth]{plots/fsratio-diff10.ps} \\caption{\\label{fig:circles} In each panel, photon events (\\emph{dots}) are either isolated (\\emph{solid blue circles}) or not (\\emph{dashed blue circles}). Random circles are either empty (\\emph{solid red}) or not(\\emph{dashed red}). In each case, 100 events and 100 random circles are shown, so the ratio, $R$, can be visualized here as the number of solid blue circles divided by the number of solid red circles. In practise, one uses a large number of random circles to reduce noise. The panels contain either no point sources (\\emph{upper left}), or 15\\% (\\emph{upper right}), 50\\% (\\emph{lower left}), or 90\\% (\\emph{lower right}) point source flux in sources located at (0.3,0.3) and (0.7,0.7). } \\end{figure*} ", "conclusions": "We have introduced a simple and easily calculable statistic that linearly traces the fraction of flux arising from diffuse emission, as opposed to unresolved point sources. The statistic is quite insensitive to even pronounced large-scale anisotropies in the diffuse emission, such as might originate from the proximity of a bright region or angular variation in the detector exposure. The linear response of this statistic to flux originating from point sources, and its smaller variance, make it superior to the two-point correlation function as a tracer of emission from unresolved point sources. The sensitivity of the statistic to point source emission naturally depends on the luminosity function of the point sources, as a sufficiently steep power law extending to sufficiently small luminosities is strictly indistinguishable from diffuse emission. However, the statistic retains discriminatory power for spectral indices up to $\\alpha \\sim 3$, with a low-luminosity cutoff corresponding to an average of 0.1 counts, and assuming all point sources with average luminosity $\\gtrsim 10$ counts are resolved and removed. Known luminosity functions for astrophysical point sources generically have shallower slopes than this limit. When the average number of events per PSF circle exceeds 1, the original form of the statistic breaks down: however, we have described a simple generalization suitable for this case, and demonstrated its efficacy. Increasing the number of counts by taking additional sky regions into account (i.e. without increasing the density of events) improves the variance by the usual $1/N_\\mathrm{event}$ Poisson factor. This statistic generalizes readily to higher dimensions; possible applications include the study of void statistics \\citep{1986ApJ...306..358F}. As an example, we have applied this statistic to Class 3 (diffuse class) photon data from the \\Fermi LAT in the angular region relevant for study of the WMAP Haze, at energies of 10-100 GeV. We find that even with rather pessimistic assumptions for the point source luminosity function, at most $\\sim 15 \\%$ of the emission in this region can be attributed to unresolved point sources with average luminosities of $0.1+$ counts / year, and the results are consistent with $100 \\%$ diffuse emission. We wish to acknowledge helpful conversations with Marc Davis, Josh Grindlay, Igor Moskalenko, Jim Peebles and Pat Slane. We thank the anonymous referee for helpful comments. TRS is supported by a Sir Keith Murdoch Fellowship from the American Australian Association. \\begin{appendix} \\begin{onecolumn}" }, "0910/0910.3097_arXiv.txt": { "abstract": "{Wide-field \\textit{Spitzer} surveys allow identification of thousands of potentially high-$z$ submillimeter galaxies (SMGs) through their bright 24\\,$\\mu$m emission and their mid-IR colors.}{We want to determine the average properties of such $z\\sim$2 \\textit{Spitzer}-selected SMGs by combining millimeter, radio, and infrared photometry for a representative IR-flux ($\\lambda_{\\rm rest}\\sim 8\\,\\mu$m) limited sample of SMG candidates.}{A complete sample of 33 sources believed to be starbursts (``5.8\\,$\\mu$m-peakers'') was selected in the (0.5\\,deg$^2$) J1046+56 field with selection criteria $F_{\\rm 24\\,\\mu m}$\\,\\textgreater\\,400\\,$\\mu$Jy, the presence of a redshifted stellar emission peak at 5.8\\,$\\mu$m, and $r^\\prime_{\\rm Vega}$\\,\\textgreater\\,23. The field, part of the SWIRE Lockman Hole field, benefits from very deep VLA/GMRT 20\\,cm, 50\\,cm, and 90\\,cm radio data (all 33 sources are detected at 50\\,cm), and deep 160\\,$\\mu$m and 70\\,$\\mu$m \\textit{Spitzer} data. The 33 sources, with photometric redshifts $\\sim1.5\\,-\\,2.5$, were observed at 1.2\\,mm with IRAM-30m/MAMBO to an rms $\\sim$0.7\\,-\\,0.8\\,mJy in most cases. Their millimeter, radio, 7-band \\textit{Spitzer}, and near-IR properties were jointly analyzed.}{ The entire sample of 33 sources has an average 1.2\\,mm flux density of $1.56 \\pm 0.22$\\,mJy and a median of 1.61\\,mJy, so the majority of the sources can be considered SMGs. Four sources have confirmed 4\\,$\\sigma$ detections, and nine were tentatively detected at the 3\\,$\\sigma$ level. Because of its 24\\,$\\mu$m selection, our sample shows systematically lower $F_{\\rm 1.2\\,mm}/F_{\\rm 24\\,\\mu m}$ flux ratios than classical SMGs, probably because of enhanced PAH emission. A median FIR SED was built by stacking images at the positions of 21 sources in the region of deepest \\textit{Spitzer} coverage. Its parameters are $T_{\\rm dust} = 37 \\pm 8$\\,K, $L_{\\rm FIR} = 2.5 \\times 10^{12}\\,L_{\\odot}$, and SFR\\,=\\,450\\,$M_{\\odot}$\\,yr$^{-1}$. The FIR-radio correlation provides another estimate of $L_{\\rm FIR}$ for each source, with an average value of $4.1 \\times 10^{12}\\,L_{\\odot}$; however, this value may be overestimated because of some AGN contribution. Most of our targets are also luminous star-forming $BzK$ galaxies which constitute a significant fraction of weak SMGs at $1.7 \\lesssim z \\lesssim 2.3.$}{\\textit{Spitzer} 24\\,$\\mu$m-selected starbursts and AGN-dominated ULIRGs can be reliably distinguished using IRAC-24\\,$\\mu$m SEDs. Such ``5.8\\,$\\mu$m-peakers'' with $F_{\\rm 24\\mu m}$\\,\\textgreater\\,400\\,$\\mu$Jy have $L_{\\rm FIR}\\,\\gtrsim 10^{12}\\,L_{\\odot}$. They are thus $z \\sim 2$ ULIRGs, and the majority may be considered SMGs. However, they have systematically lower 1.2\\,mm/24\\,$\\mu$m flux density ratios than classical SMGs, warmer dust, comparable or lower IR/mm luminosities, and higher stellar masses. About 2000\\,$-$\\,3000 ``5.8\\,$\\mu$m-peakers'' may be easily identifiable within SWIRE catalogues over 49\\,deg$^2$.} ", "introduction": "Ultra-Luminous InfraRed Galaxies (ULIRGs, with $L_{\\rm FIR} \\gtrsim 10^{12}\\,L_{\\odot}$) are the most powerful class of star-forming galaxies. For 25 years, these prominent sources and their intense starbursts have been the target of many comprehensive studies, both locally \\citep[e.g., ][]{Sand96,Lons06,Veil09} and at high redshift \\citep[e.g., ][]{Blai04,Solo05}. While local ULIRGs are relatively rare, submm/mm surveys with large bolometer arrays such as JCMT/SCUBA [James Clerk Maxwell Telescope/Submillimetre Common User Bolometer Array \\citep[]{Holl99}], APEX/LABOCA [Atacama Pathfinder Experiment/Large Apex Bolometer Camera \\citep[]{Siri09}] or IRAM/MAMBO [Institut de Radioastronomie Millim\\`etrique/Max-Planck Bolometer Array \\citep[]{Krey98}] have shown that the como\\-ving density of submillimetre galaxies (SMGs), which represent a significant class of high-redshift ($z\\sim1-4$) ULIRGs, is about a thousand times greater than that of ULIRGs in the local Universe \\citep[e.g.,][]{lefl05,Chap05}. They represent a major phase of star formation at early epochs and are also characterized by high stellar masses \\citep[e.g.,][]{Bory05}. They are thus ideal candidates to be the precursors of local massive elliptical galaxies \\citep[e.g.,][ hereafter Lo09, and references therein]{Blai02,Dye08,Lons08}. Nearly all of the enormous UV energy produced by their massive young stars is absorbed by interstellar dust and re-emitted at far-infrared wavelengths, with their far-infrared luminosity ($L_{\\rm FIR}$) able to reach $10^{13}\\,L_{\\sun}$. However, despite the considerable efforts invested in mm/submm surveys, the total number of known SMGs remains limited to several hundred, and current observational capabilities are still somewhat marginal at many wavelengths. We thus still lack comprehensive studies of SMGs and their various subclasses at all wavelengths and redshifts and in various environments. Even their star formation rates (SFRs) remain uncertain in most cases because of a lack of direct observations at the FIR/submm wavelengths of their maximum emission. The identification of large samples of SMGs is important for investigating the properties of these galaxies (SFR, luminosity, spectral energy distribution [SED], stellar mass, AGN content, spatial structure, radio and X-ray parameters, clustering, etc.) on a statistical basis, as a function of their various subclasses, redshift, and environment. This is the main goal of the wide-field submm surveys planned with SCUBA2 and \\textit{Herschel}. Although \\textit{Spitzer} generally lacks the sensitivity to detect SMGs in the far-IR, its very good sensitivity in the mid-IR allows the efficient detection of a significant fraction of SMGs in the very large area observed by its wide-field surveys, and in particular the $\\sim49\\,\\rm{deg}^2$ {\\it Spitzer} Wide-area Infrared Extragalactic (SWIRE) survey \\citep{Lons03}. From an analysis of a sample of $\\sim 100$ SMGs observed with \\textit{Spitzer}, Lo09 have estimated that SWIRE has detected more than 180 SMGs with $F_{\\rm 1.2\\,mm} > 2.5\\,$mJy per square degree at 24\\,$\\mu$m and in several IRAC bands from 3.6 to 8.0\\,$\\mu$m. However, the identification of SMGs among SWIRE sources is not straightforward, since it requires inferring FIR emission from mid-IR photometry in objects with various SEDs, especially as regards AGN versus starburst emission, and various redshifts. We have therefore undertaken a systematic study of the 1.2\\,mm emission from the best SMG candidates among \\textit{Spitzer} bright 24\\,$\\mu$m sources, focusing on $z\\,\\sim$\\,2 starburst candidates. In \\citet{Lons06} and Lo09 \\citep[see also][]{Weed06b,Farr08}, it is shown that selecting sources with a secondary maximum emission in one of the intermediate IRAC bands at 4.5 or 5.8\\,$\\mu$m provides an efficient discrimination against AGN power-law SEDs. In particular, 24\\,$\\mu$m bright ``5.8\\,$\\mu$m-peakers'' have a high probability of being dominated by a strong starburst at $z\\sim 2$, whose intense 7.7\\,$\\mu$m feature is redshifted into the 24\\,$\\mu$m band. A first 1.2\\,mm MAMBO study of a sample of $\\sim 60$ bright SWIRE sources (Lo09) has confirmed that such a selection yields a high detection rate at 1.2\\,mm and a significant average 1.2\\,mm flux density, showing that the majority of such sources are high-$z$ ULIRGs, probably at $z\\sim 2$. However, as described in Lo09, this sample was selected with the aim of trying to observe the ``5.8\\,$\\mu$m-peakers'' with the strongest mm flux over more than 10\\,deg$^{2}$. This was achieved by deriving photometric redshifts, estimating the expected 1.2\\,mm flux densities by fitting templates of various local starbursts and ULIRGs to the optical and infrared (3.6$-$24\\,$\\mu$m) bands, and selecting the candidates predicted to give the strongest mm emission. Therefore, the selection criteria of this sample were biased, especially toward the strongest 24\\,$\\mu$m sources and those in clean environments. We report here the results of an analogous MAMBO study, but of a complete 24\\,$\\mu$m-flux limited sample of all SWIRE ``5.8\\,$\\mu$m-peakers'' in a 0.5\\,deg$^{2}$ region within the SWIRE Lockman Hole field, with $F_{\\rm 24\\,\\mu m} > 400\\,\\mu$Jy and $r^\\prime_{\\rm Vega} > 23$ (see Sec.\\ 2 for a precise definition of ``5.8\\,$\\mu$m-peakers'', which of course depends on the actual SWIRE data and limits of sensitivity and accuracy). This region was selected because of the richness in multi-wavelength data, in particular the exceptionally deep radio data at 20\\,cm \\citep[VLA,][]{OwMo09}, 50\\,cm (GMRT, Owen et al. in prep.), and 90\\,cm \\citep[VLA, ][]{Owen09}. Our study aims at characterizing the average multi-wavelength properties of these sources, their dominant emission processes (starburst or AGN), their stellar masses, and their star formation rates. We adopt a standard flat cosmology: $H_{0}$=71\\,km\\,s$^{-1}$\\,Mpc$^{-1}$, $\\Omega_{M}$=0.27 and $\\Omega_{\\Lambda}$=0.73 \\citep{Sper03}. ", "conclusions": "The aim of this project was to determine the average properties of a complete 24\\,$\\mu$m flux limited sample of bright \\textit{Spitzer} sources selected to be starburst dominated at $z\\sim$2, using multi-wavelength data. The sample of 33 $z\\sim$2 SMG candidates was built with all the optically faint sources in a $\\sim$0.5\\,deg$^{2}$ area of the Lockman Hole SWIRE field meeting selection criteria based on MIPS/IRAC fluxes. These criteria are $F_{\\rm{24\\,\\mu m}}$\\,\\textgreater\\,400\\,mJy, a peak in the 5.8\\,$\\mu$m IRAC band due to redshifted 1.6\\,$\\mu$m stellar emission, and $r^{\\prime}_{\\rm{Vega}}$\\,$>$\\,23. The J1046+59 field was selected because of the availability of very deep radio observations at 20\\,cm and 90\\,cm with the VLA and at 50\\,cm with the GMRT. All sources in our sample are detected at 50\\,cm. The entire sample has an average 1.2\\,mm flux density of 1.56$\\pm$0.22\\,mJy. However, the limited sensitivity allowed only four confirmed 4\\,$\\sigma$ detections, plus nine tentative 3\\,$\\sigma$ detections. Since the average 1.2\\,mm flux density, 1.56\\,mJy, corresponds to a 850\\,$\\mu$m flux density close to 4\\,mJy, about half of the sources may be considered SMGs. However, their redshifts range from $z\\sim$1.7$-$2.3, similarly to the sample in \\citet{Huan08}, but smaller than the redshift range covered by SMGs. The sample selected here is characterized by brighter 24\\,$\\mu$m flux densities, on average, than those of SMGs, and consequently shows systematically lower $F_{\\rm{1.2\\,mm}}$/$F_{\\rm{24\\,\\mu m}}$ ratios than classical SMGs. It is quite likely that our selection favours the SMGs with the brightest 24\\,$\\mu$m flux densities, due probably to enhanced PAH emission. From stacking individual images of the sources, we are able to build the median FIR SED of our sample and estimate the corresponding $L_{\\rm{FIR}}$, SFR and $T_{\\rm{dust}}$ assuming a single temperature ``greybody'' model. The inferred values are $T_{\\rm{dust}}$\\,=\\,37$\\pm$8\\,K, $L_{\\rm{FIR}}$\\,=\\,2.5$\\times$10$^{12}$\\,$L_{\\odot}$, and SFR\\,=\\,450\\,$M_{\\odot}$\\,yr$^{-1}$. These estimates indicate that most of the sources are ULIRGs. However, estimates of $L_{\\rm{FIR}}$ for individual sources deduced from the IR-mm SED are highly uncertain due to the lack of flux measurements between 100\\,$\\mu$m and 500\\,$\\mu$m. The high quality radio data provide important complementary information on $L_{\\rm FIR}$ and the star formation rate, since the 1.2\\,mm/radio flux density ratio of the majority of individual sources is consistent with the FIR/radio correlation, which allows a derivation of $L_{\\rm{FIR}}$ and SFR from the radio flux, providing further confirmation that most of the selected sources are ULIRGs. The average value of $L_{\\rm{FIR}}$ inferred from the FIR-radio correlation is 4.1$\\times$10$^{12}$\\,$L_{\\odot}$; however, this value may be overestimated because of an AGN contribution. Stellar masses are estimated by modelling the optical-IR SED with stellar population synthesis models. They are of order a few 10$^{11}$\\,$M_{\\odot}$. Roughly scaling with the observed 5.8\\,$\\mu$m fluxes, they are similar to those of other samples of 24\\,$\\mu$m-bright, $z\\sim$2 \\textit{Spitzer} starbursts, and slightly higher than those of classical SMGs. Overall, this sample appears similar to other samples of \\textit{Spitzer} $z\\sim$2 SMGs~\\citep[Lo09;][]{Youn09} in terms of millimetre emission, $L_{\\rm FIR}$, and SFR. The complete radio detection of all sources provides a good estimate of the total star formation rate of such sources. They represent a significant fraction of all SMGs in the redshift range $z\\sim$\\,1.7-2.3 ($\\sim$10$-$15\\%). Most of these ``5.8\\,$\\mu$m-peakers'' are star-forming BzK galaxies with luminosities at the top of the luminosity distribution of sBzKs. The surface density of ``5.8\\,$\\mu$m-peakers'' has been found to be 61\\,deg$^{-2}$ by \\citet{Farr06}. This is consistent with 33 sources in the 0.49\\,deg$^{2}$ of our field. We may thus estimate that 40\\,$-$\\,60 similar ``5.8$\\mu$m-peakers'' per square degree could be identified in the full SWIRE survey (49\\,deg$^{2}$). Most of them should be $z\\sim2$ starburst ULIRGs. At least half of them may be considered to be SMGs, including a small fraction of composite obscured AGN/starburst objects. Another significant fraction may be considered as SFRGs. These results illustrate the power of deep multi-$\\lambda$ studies, especially with complete radio data, for analysing populations of powerful high-$z$ IR and submm sources. Such deep radio data are essential for disentangling starbursts and infrared-bright AGN, and for easily providing estimates of their star formation rates. We note especially the impressive complete detection of relatively weak $z\\sim$2 SMGs over 0.5\\,deg$^2$ in a single pointing of the GMRT at 610\\,MHz. As already proved by the analysis of SCUBA sources \\citep[e.g.,][]{Ivis02}, radio data are essential for identifying optical/near-IR counterparts and analysing submm surveys. This will be even more crucial for future surveys at the confusion limits of instruments like {\\it Herschel} at 300-500\\,$\\mu$m and SCUBA2 at 850\\,$\\mu$m. Even as we wait for EVLA and the new generation of SKA precursors, our results show that the GMRT at 610\\,MHz and even 325\\,MHz can already currently provide sensitivity well matched to wide {\\it Herschel} surveys. It would be interesting to explore further whether the main properties which characterize this sample, i.e. strong MIR emission, radio activity, and high stellar mass, are related. Some of these properties are likely the result of biases introduced by our selection; however, this is unlikely to be the case for all of them, especially for the radio properties. In particular, a comparison between the starburst morphology (traced by young stars, dust, PAHs or CO emission, as measured by ALMA or {\\it JWST}) and the radio size would probe whether the radio emission is produced by the starburst or by an AGN, and whether the parameters of the starburst are different from those of most classical SMGs and reveal a different star formation regime. We have several multi-wavelength observations planned or in progress for this sample to obtain better estimates of redshifts, dust temperatures, star formation rates, PAH luminosities, and AGN contributions, and thus constrain the dominant emission processes, and investigate the evolution and clustering properties of these sources." }, "0910/0910.3742_arXiv.txt": { "abstract": "We present the results of three separate searches for \\HI\\ 21-cm absorption in a total of twelve damped Lyman-$\\alpha$ absorption systems (DLAs) and sub-DLAs over the redshift range $z_{\\rm abs}=0.86-3.37$. We find no absorption in the five systems for which we obtain reasonable sensitivities and add the results to those of other recent surveys in order to investigate factors which could have an effect on the detection rate: We provide evidence that the mix of spin temperature/covering factor ratios seen at low redshift may also exist at high redshift, with a correlation between the 21-cm line strength and the total neutral hydrogen column density, indicating a roughly constant spin temperature/covering factor ratio for all of the DLAs searched. Also, by considering the geometry of a flat expanding Universe together with the projected sizes of the background radio emission regions, we find, for the detections, that the 21-cm line strength is correlated with the size of the absorber. For the non-detections it is apparent that larger absorbers (covering factors) are required in order to exhibit 21-cm absorption, particularly if these DLAs do not arise in spiral galaxies. We also suggest that the recent $z_{\\rm abs} = 2.3$ detection towards TXS 0311+430 arises in a spiral galaxy, but on the basis of a large absorption cross-section and high metallicity, rather than a low spin temperature \\citep{ykep07}. ", "introduction": "Damped Lyman-$\\alpha$ absorption systems (DLAs) are believed to be the precursors of modern day galaxies, containing at least 80\\% of the neutral gas mass density of the Universe \\citep{phw05}. As the name suggests, DLAs are identified through their heavily damped absorption features, due to the large columns of neutral hydrogen ($N_{\\rm HI}\\geq2\\times10^{20}$ \\scm) through which the background quasar is viewed. The 21-cm spin-flip transition is of interest since it traces the cool component of the gas, with the comparison of the 21-cm absorption strength to the total neutral hydrogen column giving the gas spin temperature ($T_{\\rm spin}$) for a fully absorbed emission region ($f=1$) [see Equ.~\\ref{enew}, Sect. \\ref{or}]. Searches for 21-cm absorption in DLAs exhibit a $\\approx50\\%$ detection rate, the detections occuring predominately at low redshift ($z_{\\rm abs}\\lapp1$), suggesting an increase in the $T_{\\rm spin}/f$ ratio with redshift, which could be due to higher spin temperatures and/or lower covering factors. In order to shed light on which is the predominant factor, we have undertaken searches with the Green Bank (GBT) and Giant Metrewave Radio (GMRT) Telescopes. In particular, we aim to: \\begin{enumerate} \\item Test the hypothesis of \\citet{cmp+03} that 21-cm absorption should be readily detectable at high redshift, despite the large $T_{\\rm spin}/f$ ratios, by searching for 21-cm absorption in DLAs which lie towards compact radio sources. \\item Test the hypothesis of \\citet{ctp+07} that the spin temperature does not continue to increase with redshift, by searching in previously unsearched high redshift ($z_{\\rm abs}\\gapp3.2$) DLAs. \\item Test the line strength--metallicity correlation ($T_{\\rm spin}/f$\\,--\\,[M/H] anti-correlation) of \\citet{ctp+07}, which suggests that several, currently undetected, DLAs should be readily detectable in 21-cm absorption. \\end{enumerate} We have observed a dozen DLAs and sub-DLAs and in this paper we present and discuss the results in the context of the above issues. ", "conclusions": "\\subsection{Covering factors} \\subsubsection{Angular diameter distances and other effects} \\label{copr} As mentioned in Sect. \\ref{GBT06}, \\citet{cmp+03} suggested that, contrary to the high redshift equals high spin temperature argument of \\citet{kc02}, 21-cm absorption should be readily detectable at high redshift towards compact radio sources, where the covering factor is maximised. At the time of writing up the results presented in \\citet{cw06}, this was confirmed by the detection of 21-cm absorption in the highest redshift example to date, at $z_{\\rm abs}=2.347$ towards 0438--436 (\\citealt{kse+06}), which has a background radio source size of only $0.039''$ at 5 GHz \\citep{tml+98}. Following this, since submitting the proposal for the 2006 GBT observations, 21-cm absorption has been confirmed at yet higher redshift, $z_{\\rm abs}=3.386$ towards 0201+113 (\\citealt{kcl06}, see also \\citealt{dob96,bbw97}), as well as at $z_{\\rm abs}=2.289$ towards 0311+430 \\citep{ykep07}. Again, both of these occult compact radio sources of $<0.007''$ at 1.4 GHz \\citep{sbom90}\\footnote{\\citet{kcl06} report a de-convolved source size of $17.6\\times6.6$ mas$^2$ at 328 MHz.} and $1.36''\\times0.63''$, respectively. The latter was noted by \\citet{ykep07} to be unusual in its low spin temperature of $T_{\\rm spin}\\leq140$~K at such a high redshift, although this is not unusual in regard to the finding of \\citet{cmp+03,cw06}. From Table \\ref{size} we see that five of our targets are towards radio source sizes of $\\theta_{\\rm QSO}\\lapp1''$ (recently found to be $0.07''$ for 0454+039 and $0.17''$ for 0528--250, \\citealt{klm+09}), although we have only meaningful $T_{\\rm spin}/f$ limits for three of these (0454+039, 0528--250 \\& 0758+475 -- Table~\\ref{res}). However, all of the GBT 2006 targets (Sect. \\ref{GBT06}) were selected before the arguments of \\citet{cw06} were formulated, and so all of the 2006 targets (and two of the three just mentioned), are all at high redshift, and will therefore be disadvantaged with respect to the DLA-to-QSO angular diameter distance ratios. \\begin{figure*} \\centering \\includegraphics[angle=270,scale=0.73]{2-distance-z-hist-klm+09-actual.eps} \\caption{Top: The absorber/quasar angular diameter distance ratio versus the absorption redshift. Updated from \\citet{cw06}, where the symbols are as per Fig. \\ref{Toverf} and we include only the DLAs which have been searched to $T_{\\rm spin}/f\\geq100$~K (see main text). The iso-redshift curves show how $DA_{\\rm DLA}/DA_{\\rm QSO}$ varies with absorption redshift, where $DA_{\\rm QSO}$ is for a given QSO redshift, given by the terminating value of $z_{\\rm abs}$ (throughout this paper we use $H_{0}=71$~km~s$^{-1}$~Mpc$^{-1}$, $\\Omega_{\\rm matter}=0.27$ and $\\Omega_{\\Lambda}=0.73$). That is, we show the $DA_{\\rm DLA}/DA_{\\rm QSO}$ curves for $z_{\\rm em}$ = 0.5, 1, 2, 3 \\& 4. Bottom: The radio source size versus the absorption redshift. The symbols are colour coded according to the ratio of the frequency of the image to the redshifted 1420 MHz frequency: green -- $\\nu_{\\rm size}/ \\nu_{\\rm abs}\\leq3$, orange -- $3<\\nu_{\\rm size}/ \\nu_{\\rm abs}\\leq10$ and red -- $\\nu_{\\rm size}/ \\nu_{\\rm abs} > 10$. The hatched histogram represents the detections and the bold histogram the non-detections.} \\label{2-distance-z} \\end{figure*} Showing these in Fig. \\ref{2-distance-z} (top), we see that all but one of our target DLAs (for which we have limits), are indeed at angular diameter distance ratios of $DA_{\\rm DLA}/DA_{\\rm QSO}\\approx1$, due to their high redshift selection. For a given absorber and radio source size, this will disadvantage the effective covering factor (possibly giving the observed $\\overline{[T_{\\rm spin}/{f}]_{z \\geq 1}}\\approx2\\,\\overline{[T_{\\rm spin}/{f}]_{z < 1}}$, Sect. \\ref{oftpr}), although, as mentioned above, the three new high redshift detections occult compact radio sources, as shown in the bottom panel of Fig. \\ref{2-distance-z}, where we use the source sizes compiled in \\citet{cmp+03} updated with those given in \\citet{klm+09}\\footnote{\\label{foot7}Many of the source sizes are measured at (VLBI) frequencies, which are often many times higher than that of the redshifted 21-cm line (see table 2 of \\citealt{cmp+03}), and so in Fig. \\ref{2-distance-z} (bottom) we have flagged each of the DLAs according to how many times larger than $1420/(z_{\\rm abs}+1)$ the frequency at which the source size is measured, i.e. $\\nu_{\\rm size}/\\nu_{\\rm abs}$. The recently published low frequency VLBA imaging of \\citet{klm+09} accounts for many of the ``green'' values, where $\\nu_{\\rm size}/ \\nu_{\\rm abs}\\leq3$. Since a number of these sources appear resolved, we have recalculated the extents, rather than using the deconvolved sizes quoted in \\citet{klm+09}, which are often smaller than the beam and may be unphysical (e.g. $\\sim0$ pc for 0405--331).}. However, all of our non-detections (which are predominately at $z_{\\rm abs}\\gapp2.4$), also occult similar radio source sizes. Of these, three of the sizes are not as reliable (orange symbols where $3 < \\nu_{\\rm size}/ \\nu_{\\rm abs}<10$, cf. green where $\\nu_{\\rm size}/ \\nu_{\\rm abs}\\leq3$), although as seen from Fig. \\ref{Toverf}, only three published non-detections are sufficiently sensitive to detect $T_{\\rm spin}/{f}\\gapp2000$~K at $z_{\\rm abs}\\gapp2$ (Sect. \\ref{GMRT08}). While, due to geometry effects, a high angular diameter distance will disadvantage the absorption cross-section in comparison to the same absorber placed at a low angular diameter distance, with the 21-cm searches spanning a look-back time range of 12 Gyr there are also evolutionary effects to be considered. For instance, an evolution in the morphologies of the absorbing galaxies is expected to result in different metallicities and these heavy elements can provide cooling pathways for the diffuse gas (\\citealt{dm72}). Furthermore, it is shown that the relative mix of the cold neutral medium (CNM, where $T\\sim150$ K and $n\\sim10$ \\ccm) and the warm neutral medium (WNM, $T\\sim8000$ K and $n\\sim0.2$ \\ccm) can be affected by the metallicity \\citep{whm+95}. At low redshifts, where the DLA hosts can be imaged directly, the absorbing galaxies appear to be a mix of spirals, dwarf and LSBs (e.g. \\citealt{lbbd97,rnt+03,cl03}) and, in general, large galaxies appear to be more populous at low redshift with smaller morphologies becoming more common at high redshifts \\citep{bmce00,lf03}. Therefore, in the more compact absorbing galaxies, where the metallicities are lower than in the larger spirals (see Sect. \\ref{lsmc}), we may expect the gas to be more dominated by the WNM and, as shown by \\citet{ctp+07}, $T_{\\rm spin}/{f}$ is anti-correlated with the metallicity (see Sect. \\ref{lsmc}). However, although the metallicity is itself anti-correlated with the look-back time (\\citealt{pgw+03,rphw05}), no [M/H]--$z_{\\rm abs}$ correlation is seen for the 21-cm searched sample \\citep{ctp+07}, and, as seen in Fig.~\\ref{2-distance-z}, very few morphologies are known for these at $z_{\\rm abs}\\gapp1$. Furthermore, the $T_{\\rm spin}/{f}$\\,--\\,[M/H] anti-correlation is consistent with {\\em either} the spin temperature being lower or the covering factor being larger at high metallicities, with the larger galaxies likely to have high values of $f$ as well as [M/H]. Although the $T_{\\rm spin}/{f}$ degeneracy is unbreakable, it is possible that both factors contribute and that the values of [M/H], $T_{\\rm spin}$ and ${f}$ are in fact interwoven \\citep{ctp+07}. In addition to decreasing metallicities, the increase in the background ultra-violet flux will increase the ionisation fraction of the gas \\citep{bdo88}, in addition to raising its spin temperature \\citep{fie59,be69}, an effect also seen in associated systems, where 21-cm absorption has never yet been detected in the host galaxy of a quasar with $L_{\\rm UV}\\gapp10^{23}$ W Hz$^{-1}$ \\citep{cww+08}. Therefore, as well as always having large angular diameter distance ratios and lower metallicities, high redshift absorption systems will generally be subject to higher ultra-violet fluxes, although for intervening absorbers the effect will not be as severe as for associated absorbers. Nevertheless, this could be the cause of many of the $T_{\\rm spin}/{f}$ lower limits at $z_{\\rm abs}\\gapp2.5$ (Fig. \\ref{Toverf}), although, again, such heated gas could have a significantly lower 21-cm absorption cross-section as well as a higher temperature. \\subsubsection{Absorber extents} \\label{aex} In order to consider the radio source size in conjunction with the ratio of the angular diameter distances, we calculate the extent of absorber required to fully cover the radio source (Fig.~\\ref{dla_schem}) and show this against the normalised line \\begin{figure} \\vspace{3.0cm} \\special{psfile=dla_schem.eps hoffset=-20 voffset=0 hscale=45 vscale=45 angle=0} \\caption{In the small angle approximation $\\theta = d_{\\rm abs}/DA_{\\rm DLA} = d_{\\rm QSO}/DA_{\\rm QSO}$, giving $d_{\\rm abs (min)}=\\theta_{\\rm QSO}.\\,DA_{\\rm DLA}$ when the absorber just covers the background radio source size.} \\label{dla_schem} \\end{figure} strength, $1.823\\times10^{18}\\,\\int\\tau\\,dv/N_{\\rm HI}= f/T_{\\rm spin}$ (Fig. \\ref{size-z}). \\begin{figure} \\vspace{6.5cm} \\special{psfile=size-foverT-klm+09-actual.ps hoffset=-20 voffset=230 hscale=75 vscale=75 angle=270} \\caption{The minimum extent of the absorber required to cover the radio source versus the normalised line strength. The symbols are as per Fig. \\ref{2-distance-z} (bottom). The line shows the least-squares fit to the detections. The upper limits in the ordinate are due to upper limits in the background radio source sizes.} \\label{size-z} \\end{figure} From this we see a correlation between the extent of the absorber and the 21-cm absorption strength for the detections. For the 18 detections this is significant at $2.7\\sigma$, falling to $2.2\\sigma$ for the 12 detections with $\\nu_{\\rm size}/ \\nu_{\\rm abs}\\leq3$ (``green''), with the non-detections exhibiting no correlation whatsoever (bringing the significance for the whole 39 strong sample down to $1.4\\sigma$)\\footnote{Using the deconvolved values of \\citet{klm+09}, rather than those measured (see footnote \\ref{foot7}), gives $2.5\\,,1.9$ \\& $1.3\\sigma$, respectively.}. The fit in Fig. \\ref{size-z} suggests that the non-detections which have been searched sufficiently deeply, (i.e. neglecting the two with $f/T_{\\rm spin}\\gapp0.01$), may require large absorbers ($\\gapp10$~kpc scales) in order to achieve effective covering factors. As is seen from the figure, several detections also occupy this regime, although these tend to be spiral galaxies. Therefore, like \\citet{kc02}, and as discussed in \\citet{cmp+03}, the majority of non-detections (which are at high redshift and have unknown morphologies) may be non-spirals. Although, unlike \\citet{kc02}, we again suggest that may be their small sizes, in addition to/rather than high mean spin temperatures, which may be responsible for their weak line strengths. The covering factor can be defined\\footnote{It can also be estimated as the ratio of the compact unresolved component's flux to the total radio flux \\citep{bw83,klm+09}. However, even if high resolution radio images at the redshifted 21-cm frequencies are available, this method gives no information on the extent of the absorber (or how well it covers the emission). Furthermore, the high angular resolution images are of the continuum only and so do not give any information about the depth of the line when the extended continuum emission is resolved out.} as $f\\equiv [d_{\\rm abs}/(\\theta_{\\rm QSO}.\\,DA_{\\rm DLA})]^2$ \\citep{cw06}, meaning that, in general\\footnote{That is, where $d_{\\rm abs}$ is the actual extent of the absorber, rather than the extent minimum required to cover $\\theta_{\\rm QSO}$.}, the ordinate in Fig. \\ref{size-z} is equivalent to $d_{\\rm abs}/\\sqrt{f}$. This therefore suggests, at least for the detections, that the line strength increases with the size of the 21-cm absorption region, a trend which is apparent, even when the absorber size is weighed down by the $\\sqrt{f}$ factor. and, in general, from the normalised line strength, \\begin{equation} 1.823\\times10^{18}\\,\\frac{\\int\\tau\\,dv}{N_{\\rm HI}}= \\frac{f}{T_{\\rm spin}} \\propto\\theta_{\\rm QSO}.\\,DA_{\\rm DLA} = \\frac{d_{\\rm abs}}{\\sqrt{f}}. \\end{equation} That is, $d_{\\rm abs}\\propto f^{3/2}/T_{\\rm spin}$, which makes sense in terms of both covering factor and spin temperature -- a larger absorber implies a larger covering factor, as well as the possibility of a lower spin temperature. Presuming that this is not dominated by the numerator, this {\\em could} suggest that the larger absorbing galaxies are subject to lower spin temperatures (\\citealt{kc02} and references therein). However, one interpretation of the distribution of the non-detections along the ordinate in Fig. \\ref{size-z} is that for a given absorption cross-section (i.e. assuming the same $d_{\\rm abs}$ for all the DLAs), the covering factor is generally lower than for a number of the detections. Note that, in addition to geometry effects, it is possible that these lower covering factors could also be caused by lower beam filling factors, introduced by structure (``clumpiness'') in the absorbing gas, which may be behind the 21-cm variability seen at $z_{\\rm abs}=0.313$ towards 1127--145 \\citep{kc01} and at $z_{\\rm abs}=3.386$ towards 0201+113 \\citep{dob96,bbw97,kcl06}. \\citet{gsp+09a} also note from their large survey of Mg{\\sc \\,ii} absorption systems that the decreasing number density of 21-cm absorbers with redshift runs counter to that of the Lyman-\\AL\\ and ${\\rm W}_{\\rm r}^{\\lambda2796}>1$ \\AA\\ Mg{\\sc \\,ii} absorbers \\citep{rtn05}. This is attributed to a significant evolution in either the CNM filling factor or the covering factor, the former of which could be the result of a relatively extensive WNM, thus indicating a higher mean harmonic spin temperature \\citep{klm+09}, although the larger fractions of warm gas would not exclude the presence of cold clumps. Either or both of these scenarios would be tied up in the $T_{\\rm spin}/f$ degeneracy. Nevertheless, 21-cm detection rates are significantly lower at high redshift and the filling factor of the CNM, through either its effect on $T_{\\rm spin}$ or $f$, could be a major contributor to this. \\subsection{21-cm line strength--metallicity correlation} \\label{lsmc} Addressing the 21-cm line strength--metallicity correlation ($T_{\\rm spin}/f$\\,--\\,[M/H] anti-correlation) of \\citet{ctp+07}, the focus of the 2008 GBT observations (Sect. \\ref{GBT08}), it appears that we should have detected absorption in at least one of the two DLAs for which reasonable sensitivities were reached (Fig. \\ref{foverT-m}). \\begin{figure} \\vspace{6.5cm} \\special{psfile=foverT-m.ps hoffset=-20 voffset=230 hscale=75 vscale=75 angle=270} \\caption{The normalised line strength ($\\equiv f/T_{\\rm spin}$) versus the metallicity for the DLAs searched in 21-cm absorption, where the symbols are as per Fig.~\\ref{Toverf}. The significance of the correlation is given for both the whole sample and the detections only and the least-squares fit for the detections shown.} \\label{foverT-m} \\end{figure} Regarding the best improved limit\\footnote{By a factor of $\\approx7$, cf. \\citealt{cld+96}. See footnote \\ref{cld}.}, 21-cm absorption may not have been detected towards 0528--250, on the grounds that, with an absorption redshift very close to that of the emission redshift (both at $z=2.8110$, \\citealt{mff04} and references therein), the absorber is possibly located in the host of the background quasar. We may therefore expect the neutral gas to be subject to excitation effects seen in other high redshift 21-cm absorption searches \\citep{cww+08}. However, the presence of H$_2$ in the absorber suggests that the gas is relatively cool with $T_{\\rm kin}\\approx 110 - 150$ K (\\citealt{spl+05}) and so this is unlikely. In any case, referring to Fig.~\\ref{foverT-m}, the limit is not overly low in comparison to 0438--436 (the detection at [M/H]$\\,=-0.68$, $f/T_{\\rm spin}=6.3\\times10^{-4}$ [K$^{-1}$]) and there is also the possibility that we have missed the line, with our limit only being good over the redshift range $z\\approx 2.806-2.812$ (Sect. \\ref{GBT08}). Assuming this limit is reliable and including 0311+430, strengthens the correlation (cf. \\citealt{ctp+07}). \\citet{ykep07} comment that the $z_{\\rm abs}=2.289$ absorber towards 0311+430 has an unusually high metallicity, although according to Fig. \\ref{foverT-m}, this has some leeway before the metallicity becomes atypically high. From its grouping in the line strength--metallicity plot, we suggest that the absorption is due to a spiral galaxy, as do \\citet{ykep07} on the basis of its low ``spin temperature''. Support for this assertion is given by its grouping with the spirals in Fig. \\ref{size-z} (the ``unknown'' at $7\\times10^{-3}$ K$^{-1}$, 5 kpc), although this is on the grounds of a large absorption cross-section, rather than the spin temperature." }, "0910/0910.5813_arXiv.txt": { "abstract": "{}{The detection of rotational transitions of the AlO radical at millimeter wavelengths from an astronomical source has recently been reported. In view of this, rotational transitions in the ground \\textit{X}\\textsuperscript{2}$\\Sigma$\\textsuperscript{+} state of AlO have been reinvestigated. } {Comparisons between Fourier transform and microwave data indicate a discrepancy regarding the derived value of $\\gamma$$_{D}$ in the v = 0 level of the ground state. This discrepancy is discussed in the light of comparisons between experimental data and synthesized rotational spectra in the v = 0, 1 and 2 levels of \\textit{X}\\textsuperscript{2}$\\Sigma$\\textsuperscript{+}.} {A list of calculated rotational lines in v = 0, 1 and 2 of the ground state up to \\textit{N}' = 11 is presented which should aid astronomers in analysis and interpretation of observed AlO data and also facilitate future searches for this radical.}{} ", "introduction": "\\paragraph{ The ground \\textit{X}\\textsuperscript{2}$\\Sigma$\\textsuperscript{+} state of the AlO radical has been studied with microwave spectroscopy in the v = 0, 1 and 2 levels. T\\\"{o}rring et al. (1989) recorded the \\textit{N} = 1 $\\rightarrow$2 transition near 76 GHz. In their Table 1, frequencies for several $\\Delta$\\textit{F} = $\\Delta$\\textit{N} and $\\Delta$\\textit{F} $\\neq$ $\\Delta$\\textit{N} transitions are shown. Yamada et al. (1990) and Goto et al. (1994) recorded several $\\Delta$\\textit{F} = $\\Delta$\\textit{N} rotational transitions, resulting in a set of accurate molecular constants, including $\\gamma$ and $\\gamma$$_{D}$. } \\paragraph{ Launila et al. (1994) have performed a Fourier transform study of the \\textit{A}\\textsuperscript{2}$\\Pi$$_{i}$\\textbf{$\\rightarrow$}\\textit{X}\\textsuperscript{2}$\\Sigma$\\textsuperscript{+} transition of AlO in the 2 $\\mu$m region. In that work, some discrepancies were pointed out regarding their derived $\\gamma$$_{D}$ values, as compared to those found in the microwave work. While Yamada et al. (1990) had given a positive value for $\\gamma$$_{D}$ for v = 0, the sign was in fact found to be negative in the light of high-\\textit{N} data of Launila et al. (1994). One of the aims of the present paper is to reinvestigate this discrepancy more closely. The work by Goto et al. (1994), dealing with the v = 1,2 vibrational levels of the ground state of AlO, does not show the same discrepancy, however. } \\paragraph{In a theoretical work, Ito et al. (1994) have discussed and explained the observed vibrational anomalies in the spin-rotation constants of the ground state in the light of spin-orbit interaction with the \\textit{A}\\textsuperscript{2}$\\Pi$$_{i}$ and \\textit{C}\\textsuperscript{2}$\\Pi$ states.} \\paragraph{Recently, Tenenbaum and Ziurys (2009) reported three rotational transitions of AlO from the supergiant star VY Canis Majoris. In order to facilitate future millimeter-wave search for rotational transitions of AlO, tables containing expected frequencies in the v = 0, 1 and 2 levels of the ground state are useful. In the present work, such tables are presented.} ", "conclusions": "\\paragraph{The determination of the correct constants for a molecule/radical has its own intrinsic value. Additionally, the present study is also relevant in an astronomical context. Tenenbaum and Ziurys (2009) have recently made the first radio/mm detection of the AlO (\\textit{X}\\textsuperscript{2}$\\Sigma$\\textsuperscript{+}) radical toward the envelope of the oxygen rich supergiant star VY Canis Majoris (VY CMa). They observed the \\textit{N} = 7$\\rightarrow$6 and 6$\\rightarrow$5 rotational transitions of AlO at 268 and 230 GHz and the N = 4\\textbf{$\\rightarrow$}3 line at 153 GHz. While their search for the \\textit{N} = 7$\\rightarrow$6 hyperfine transitions was based on direct laboratory measured frequencies by Yamada et al (1990), the search for the \\textit{N} = 6$\\rightarrow$5 and \\textit{N} = 4$\\rightarrow$3 transitions was based on frequencies calculated from spectroscopic constants of those authors. As has been pointed out, an error is present in one of these constants ($\\gamma$$_{D}$ in the v = 0 level of the ground state), which leads to deviations between the calculated and true frequencies. While these deviations are small and may not affect the final result of a line search too seriously, it is still desirable to have accurate frequencies to facilitate future millimeter-wave searches for rotational transitions of AlO. It is planned for several such searches to be taken place in the near future as a consequence of the recent detection in VY CMa. AlO is slowly emerging as a molecule which could attract a fair deal of interest among astronomers. In the VY CMa detection for example it is shown how its study could lead to a better understanding of the gas-phase refractory chemistry in oxygen-rich envelopes. AlO has also been proposed to be a potential molecule in the formation of alumina, which is one of the earliest and most vital dust condensates in oxygen rich circumstellar environments (Banerjee et al. 2007). The mineralogical dust condensation sequence and processes involved are issues of considerable interest to astronomers. The AlO radical also received considerable attention after the strong detection of several \\textit{A}$\\rightarrow$\\textit{X} bands in the near-infrared (1 - 2.5 microns) in the eruptive variables V4332 Sgr and V838 Mon (Banerjee et al. 2003; Evans et al 2003). Other IR detections of AlO include IRAS 08182-6000 and IRAS 18530+0817 (Walker et al. 1997). In the optical, the \\textit{B}$\\rightarrow$\\textit{X} bands have been prominently detected in U Equulei (Barnbaum et al. 1996) and in several cool stars and Mira variables including Mira itself (Keenan et al. 1969; Garrison 1997).} \\paragraph{The line lists presented here should also aid in analysis of mm line data for aspects related to kinematics. Since the rotational levels of AlO species are split by both fine and hyperfine interactions, each rotational transition consists of several closely spaced hyperfine components. Figure 2 exemplifies this. If line broadening is small and the spectral resolution of the observations is adequate to resolve these components, then different components will yield different Doppler velocities if the corresponding reference or rest frequencies are in error. This could lead to ambiguity in interpreting the data. Even if the hyperfine components are not distinctly resolved but rather blended to give a composite line profile (as in the observed profiles in VY CMa), modelling of such composite profiles using wrong rest frequencies could lead to errors in estimating the composite line centre, half width of the composite profile and half-widths of the individual hyperfine components. Proper estimates of such kinematic parameters are important as they help determine the size and site of origin of a molecular species. The detailed discussion by Tenenbaum and Ziurys (2009) in the case of VY CMa, which has three distinctly different kinematic flows in the system, illustrates this.}" }, "0910/0910.3233_arXiv.txt": { "abstract": "The detection of upward propagating internal gravity waves {\\glite at the base of} the Sun's chromosphere has recently been reported by Straus et al., who postulated that these may efficiently couple to Alfv\\'en waves in magnetic regions. This may be important in transporting energy to higher levels. Here we explore the propagation, reflection and mode conversion of linear gravity waves in a VAL C atmosphere, and find that even weak magnetic fields usually reflect gravity waves back downward as slow magnetoacoustic waves well before they reach the Alfv\\'en/acoustic equipartition height at which mode conversion might occur. However, for certain highly inclined magnetic field orientations in which the gravity waves manage to penetrate near or through the equipartition level, there can be substantial conversion to either or both upgoing Alfv\\'en and acoustic waves. Wave energy fluxes comparable to the chromospheric radiative losses are expected. ", "introduction": "\\label{intro} Using the Interferometric BIdimensional Spectrometer (IBIS) and the Echelle Spectrograph on the Dunn Solar Telescope (DST) of the Sacramento Peak National Solar Observatory, and the Michelson Doppler Imager (MDI) on the Solar and Heliospheric Observatory (SOHO), \\cite{straus08} have identified upward propagating\\footnote{The group velocity and hence energy flux is upward. As expected of gravity waves, the phase velocity is downward.} gravity waves with frequencies between 0.7 mHz and 2.1 mHz in weak magnetic field regions of the solar atmosphere, and found that their energy flux is an order of magnitude larger than co-spatial acoustic waves, and comparable to the expected quiet-sun chromospheric losses of around 4.3 $\\rm kW\\,m^{-2}$. However, the gravity waves were found to be significantly suppressed in stronger field regions. Nevertheless, in light of recent identification of ubiquitous Alfv\\'en waves in the corona \\citep{pontieu07,tomczyk07}, they postulate that when these gravity waves enter magnetic regions they may efficiently couple to Alfv\\'en waves, perhaps contributing to the observed coronal wave flux. Furthermore, \\cite{jess09} have directly identified torsional Alfv\\'en waves in H$\\alpha$ bright-point groups at frequencies as low as 1.4 mHz, which places them at least partially in the gravity wave regime if there is any coupling. In this paper, we explore the propagation, reflection, and mode conversion of atmospheric gravity waves of around 1 mHz in frequency using both dispersion relations and numerical solution of the governing differential equations in simple atmospheric models with uniform inclined magnetic field. The imposed fields of 10 to 100 Gauss are weak in the photospheric context, but become dominant at greater heights as the plasma $\\beta$ (the ratio of plasma to magnetic pressure) falls below unity due to density stratification. We therefore have a situation where low frequency waves are essentially gravity waves at low altitudes, but become magnetically dominated at higher levels in the chromosphere. The central question is: \\emph{What happens to upward propagating gravity waves as they enter regions where magnetic forces become significant? Does the magnetic field help or hinder propagation through the chromosphere?} The answer is: both, depending on magnetic field orientation. ", "conclusions": "\\label{conclusions} The main conclusions we draw from our analyses are: \\renewcommand{\\labelenumi}{\\arabic{enumi}.} \\begin{enumerate} \\item Even very weak magnetic fields very effectively reflect gravity waves back downward as slow magneto-acoustic waves. This typically happens well below the $a=c$ equipartition level. \\item However, at very large magnetic field inclinations, typically around $80^\\circ$ or more depending on frequency (see Fig.~\\ref{fig:Ffreqs}), substantial mode conversion from gravity waves to either field-guided acoustic waves (for small $\\phi$) or Alfv\\'en waves ($20^\\circ\\la \\phi \\la 70^\\circ$) occurs, and these waves continue to propagate upward along the field lines. The amount of energy they carry is potentially significant for the upper chromosphere. \\item Wave energy fluxes reaching the top of our model are very sensitive to magnetic field direction, but quite insensitive to magnetic field strength, at least in the range 10 G -- 100 G. \\item The dispersion diagrams give a simple, easy and quite accurate picture of the behaviour of gravity waves in a magneto-atmosphere, though with the caveat that tunnelling can sometimes occur between branches. \\end{enumerate} In simple terms, we conclude that atmospheric gravity waves are very effectively suppressed by even very weak magnetic field, \\emph{unless} that field is highly inclined and the attack angle fine, in which case it opens a window to the upper atmosphere that allows the gravity waves to propagate through in a different guise. This is closely related to the ``magnetic portals'' of \\cite{jeff06} which rely on the ramp effect to allow acoustic waves below the acoustic cutoff frequency to still propagate upward in low-$\\beta$ inclined magnetic field, but in the case of low-frequency gravity waves which are already propagating, it gives them the opportunity to convert to propagating acoustic waves around the $a=c$ level. Because of their low frequency though, very substantial inclination is required to open these windows, but no more than is characteristic of chromospheric canopy. {\\hilite There are consequences for recent and future observations of solar atmospheric oscillations. The surprising extent to which even very weak vertical or moderately inclined magnetic field inhibits gravity waves by causing them to quickly reflect as slow magneto-acoustic waves perhaps explains ``significantly suppressed atmospheric gravity waves at locations of magnetic flux'' found by \\cite{straus08}. The ubiquity of near-horizontal field in the low solar atmosphere discovered recently with \\emph{Hinode} \\citep{lites08} however raises the possibility that low frequency gravity waves may efficiently couple to Alfv\\'en waves that continue to propagate vertically into the corona, contributing to the vast sea of waving field lines now known to exist there \\citep{pontieu07,tomczyk07}. Although \\citeauthor{tomczyk07} detect a peak in velocity power of these Alfv\\'en oscillations at around 3.5 mHz, (presumably driven by the Sun's internal normal modes), the power spectrum continues to rise with decreasing frequency till at least 1 mHz, well inside the gravity wave regime at photospheric level. It is tempting to postulate that gravity waves may be the vector of this wave energy at low levels and that it may convert to Alfv\\'en waves around the acoustic/Alfv\\'enic equipartition level in highly inclined field regions. Further observational work, ideally at multiple heights, is warranted to more fully explore these possibilities. } It should be emphasised though that our analysis is entirely linear. Acoustic waves are likely to shock before reaching the upper chromosphere. However, Alfv\\'en waves do not suffer this fate. Our models are also adiabatic, which is not a good representation of the chromosphere, {\\hilite though the detections of \\cite{straus08} suggest that {\\glite atmospheric} radiative losses do not completely suppress gravity waves, at least at the {\\glite photospheric} altitudes sampled by IBIS} {\\glite and MDI}. The adiabatic assumption will be relaxed in future work." }, "0910/0910.3834_arXiv.txt": { "abstract": "The bulk motion of galaxies induced by the growth of cosmic structure offers a rare opportunity to test the validity of general relativity across cosmological scales. However, modified gravity can be degenerate in its effect with the unknown values of cosmological parameters. More seriously, even the `observed' value of the RSD (redshift-space distortions) used to measure the fluctuation growth rate depends on the assumed cosmological parameters (the Alcock-Paczynski effect). We give a full analysis of these issues, showing how to combine RSD with BAO (baryon acoustic oscillations) and CMB (Cosmic Microwave Background) data, in order to obtain joint constraints on deviations from general relativity and on the equation of state of dark energy whilst allowing for factors such as non-zero curvature. In particular we note that the evolution of $\\Omega_m(z)$, along with the Alcock-Paczynski effect, produces a degeneracy between the equation of state $w$ and the modified growth parameter $\\gamma$. Typically, the total marginalized error on either of these parameters will be larger by a factor $\\simeq 2$ compared to the conditional error where one or other is held fixed. We argue that future missions should be judged by their Figure of Merit as defined in the $w_p - \\gamma$ plane, and note that the inclusion of spatial curvature can degrade this value by an order of magnitude. ", "introduction": "Models of dark energy leave a characteristic signature embedded in both the cosmic expansion and structure formation histories. Recent observational progress has been made with the former, due to its relative ease of measurement, leading to a measurement of the dark energy equation of state parameter $w\\equiv P/\\rho c^2$ with better than 10\\% precision \\cite{2008arXiv0803.0547K,2009arXiv0907.1660P}. This work is geometrical, and so probes dark energy only through its influence on the evolving expansion rate of the Universe. It is thus possible that dark energy may be an illusion, indicating the need to revise general relativity and thus also the Friedmann equation. In either case, the phenomenological dark energy term may well differ from a cosmological constant ($w=-1$), and may change its equation of state with redshift. These possible degrees of freedom need to be allowed for before we can claim any evidence for a deviation from general relativity. This paper thus considers how we can make simultaneous measurements of the properties of dark energy and of modified gravity. A number of probes are capable of measuring $w$ via its influence on the redshift-distance relation. This measurement alone is effectively completely degenerate with a modification of gravity on the scale of the Hubble radius. But for many models, the Mpc scales of galaxy clustering may be affected in a different way; the growth rate of density fluctuations has thus emerged as a key means of breaking this degeneracy between gravity and dark energy \\cite{2005PhRvD..72d3529L}\\cite{2008Natur.451..541G}. It is rather more difficult to study the growth rate, due to uncertainty in the behaviour of galaxy bias, but there are currently two promising avenues available for future exploration. Weak gravitational lensing provides a direct measurement of the dark matter distribution, and its evolution with redshift. It can also probe broader aspects of modified gravity, particularly the balance between perturbations to the time and space parts of the metric \\cite{beanism,2008PhRvD..78f3503J}. The focus of the present work will be the alternative technique, known as redshift-space distortions (RSD), which exploit the relationship between the large-scale coherent velocities of galaxies and the growth rate of perturbations. In real space, we expect the clustering of galaxies to be statistically isotropic. However, in redshift space the line-of-sight component of a galaxy's peculiar velocity breaks this symmetry. Inside a virialized cluster of galaxies, the orbital velocity dispersion scatters galaxy redshifts, creating the `Fingers of God', and thereby erasing spatial information on small scales. Across larger scales, galaxies coherently fall out of voids and into overdense regions, considerably amplifying the power in redshift space. These two effects are often treated independently, although a more complex model is required to attain a higher degree of precision \\cite{scocciz}. For the present purpose, the large-scale effect is the aspect of interest, since continuity relates coherent peculiar velocities directly to the growth rate of density fluctuations. Observations to date have led to estimates of the growth rate at various redshifts, although not yet at a useful level of precision \\cite{2002MNRAS.332..311H,2005MNRAS.361..879D,2007MNRAS.381..573R,2008Natur.451..541G}. Future surveys are likely to cover orders of magnitude larger volumes, thereby delivering the precision needed to discriminate interesting models of modified gravity. But we shall see that, when approaching this target, it may no longer be appropriate to make the simplifying assumptions adopted to date. In \\S\\ref{sec:sig} we review the process of determining the growth rate from redshift distortions, before constructing a Fisher matrix. Our results are presented in \\S\\ref{sec:results}, while in \\S\\ref{sec:merit} we consider the implications for the proposed dark energy Figure of Merit. ", "conclusions": "By relaxing the common assumption of a fixed background cosmology, we have highlighted some of the difficulties encountered when attempting to study gravity via the bulk motion of galaxies. Rather than a pure probe of structure, redshift distortions also comprise a geometric component. This enters at the stage of converting the true observables, angles and redshifts, into distances and Fourier modes. Furthermore, when determining the growth index $\\gamma$ it is essential that its corresponding radix $\\Omega_m(z)$ is well determined. With these two factors in mind, it appears unlikely that the galaxy power spectrum alone could provide conclusive evidence against General Relativity. To converge on the true underlying cosmology, iterating over a value for $\\Omega_m$ has proved adequate for current data. However with the greater degrees of freedom required to test relativity ($w_0, w_a, \\Omega_k$), the available volume of parameter space appears too great. Fortunately future data will inevitably be accompanied by improved measurements of the baryon acoustic oscillations. Ironically the squashing effect that empowers the BAO is the very same Alcock-Paczynski effect that confounds the redshift distortions. One concern in the formalism may be the assumption of scale-independence for both the growth and bias. More physically motivated forms of modified gravity, such as $f(R)$ models, \\cite{2007PhRvD..76j4043H}, lead to rather different scale-dependent growth factors. However, as highlighted in \\cite{2007PhRvD..76j4043H}, such models also generate very prominent deviations on intermediate scales, which would become more immediately apparent. Nevertheless, neglect of these issues is more likely to lead to a bias in the results of analyses that assume scale-independent effects, rather than changing their statistical precision. In this work, we have concentrated on the latter aspect, and our main conclusion is that the parameters $\\gamma$ and $w_p$ will generally be strongly anti-correlated. We therefore suggest that a natural Figure of Merit for future experiments in fundamental cosmology should be the reciprocal of the area of the error contour in the $\\gamma - w_p$ plane. \\noindent{\\bf Acknowledgements} \\\\ We thank Luigi Guzzo for many helpful comments on an earlier draft of this paper, and also Thomas Kitching and Will Percival for several productive discussions. FS was supported by an STFC Rolling Grant." }, "0910/0910.0837_arXiv.txt": { "abstract": "We examine the star formation rates (SFRs) of galaxies in a redshift slice encompassing the $z = 0.834$ cluster \\rxjfull. We used a low-dispersion prism in the Inamori Magellan Areal Camera and Spectrograph (IMACS) to identify galaxies with $z_{\\rm AB} < 23.3$~mag in diverse environments around the cluster out to projected distances of $\\sim 8$~Mpc from the cluster center. We utilize a mass-limited sample (\\mcutrange) of \\nmcut\\ galaxies that were imaged by Spitzer MIPS at 24~\\micron\\ to derive SFRs and study the dependence of specific SFR (SSFR) on stellar mass and environment. We find that the SFR and SSFR show a strong decrease with increasing local density, similar to the relation at $z \\sim 0$. Our result contrasts with other work at $z \\sim 1$ that find the SFR-density trend to reverse for luminosity-limited samples. These other results appear to be driven by star-formation in lower mass systems ($M \\sim 10^{10}$~\\msun). Our results imply that the processes that shut down star-formation are present in groups and other dense regions in the field. Our data also suggest that the lower SFRs of galaxies in higher density environments may reflect a change in the ratio of star-forming to non-star-forming galaxies, rather than a change in SFRs. As a consequence, the SFRs of star-forming galaxies, in environments ranging from small groups to clusters, appear to be similar and largely unaffected by the local processes that truncate star-formation at $z \\sim 0.8$. ", "introduction": "Pioneering studies of the impact of environment on galaxy properties have found higher density environments in the local universe to be dominated by elliptical and S0 galaxies \\citep{dressler1980}. Because of the strong correlation of Hubble type with star formation rate \\citep[SFR,][]{kennicutt1998}, a correlation between SFR and local galaxy density is expected and has been seen in more recent studies of the local universe \\citep{gomez2003,balogh2004}. \\citet{kauffmann2004} examined the SDSS sample of \\citet{brinchmann2004} and found a strong dependence of SFR on local galaxy density and stellar mass at $z \\sim 0$. At a fixed stellar mass, they found the specific SFRs (SSFRs), and therefore SFRs, of galaxies to be lower in higher density environments. At $z \\sim 1$, the morphology-density relation (MDR) follows a similar trend to that found in the local universe for mass-limited samples \\citep{holden2007,vanderwel2007b}. In contrast, recent analyses of the field at $z \\sim 1$ suggest that the SFR-density trend was the {\\em reverse} at earlier times, where galaxies at higher densities display {\\em higher} SFRs than galaxies at lower densities \\citep{elbaz2007,cooper2008}. In this Letter, we explore the SFRs of galaxies in a redshift slice that includes the $z = 0.834$ galaxy cluster \\rxjfull\\ (hereafter \\rxj). Our wide-field spectroscopic survey extends to a projected distance of $\\sim 8$~Mpc from the cluster center, resulting in a large number of galaxies that span a much broader range of environments at this epoch than is found in other work. Our aim is to determine the form of the SSFR/SFR-density relation at $z \\sim 0.8$ and compare to the results discussed above at these redshifts and at $z \\sim 0$. We assume a cosmology with $H_0 = 70$~\\kmps~\\pmpc, $\\Omega_M = 0.3$, and $\\Omega_{\\Lambda} = 0.7$. Stellar masses and SFRs are based on a Chabrier IMF \\citep{chabrier2003}. ", "conclusions": "Recent work at $z \\sim 1$ indicates that for galaxies in the mass range studied here, several relations follow trends similar to those found at $z \\sim 0$. For example, \\citet{holden2007} and \\citet{vanderwel2007b} found a strong MDR for mass-limited samples in both the field and clusters at $z \\sim 1$. Thus, it should not be surprising that we see a similar trend for the SSFR/SFR-density relation at $z \\sim 0.8$ as we see at $z \\sim 0$ (although with a different normalization). Interestingly, some recent studies of galaxies at low-redshift find evidence for enhanced levels of dust-obscured SF at densities slightly above typical field densities, although they also generally find an overall trend of decreasing SF activity at higher densities \\citep{gallazzi2009,wolf2009,haines2009b}. Our result of a declining SFR in higher density environments appears universal at $z \\sim 0.8$ and not confined to a cluster environment, much like the MDR and SFR-density relations in the local universe. As seen in Figure~\\ref{stack_mips_density}, cluster galaxies dominate the highest density bin, but much of the decrease in SSFR occurs in lower density field bins. In addition, after removing galaxies within $\\sim 2.5$ times the virial radius of the cluster, we continue to find the SSFR to decrease with density, including in the remaining high density regions represented by several groups at projected distances of $\\sim 3-5$~Mpc from the cluster. This decrease in SSFR occurs over a similar range of densities in which \\citet{patel2009} found an increase in the red galaxy fraction, possibly linking the end of SF and buildup of red-sequence galaxies in environments that reach into the dense regions of the field. In contrast to our work, \\citet{elbaz2007} and \\citet{cooper2008} found very different results for the SFR-density relation. At $0.8 < z < 1.2$, \\citet{elbaz2007} found a factor of $\\sim 6$ spread in SFRs. However, they found galaxies at higher densities to have {\\em higher} mean SFRs up to a critical density, above which the SFR declined. Much of the reversed SFR-density trend in \\citet{elbaz2007} is driven by a peak in the SFR in a narrow projected density range of $\\sim 0.1$~dex ($3 < \\Sigma~(\\rm Mpc^{-2}) < 4$). Likewise, \\citet{cooper2008} also found the SFRs of galaxies to increase in higher density environments at $0.75 < z < 1.05$, although their observed spread in SFRs was less than a factor of $\\sim 1.5$. While neither of these two surveys contain a cluster, both sample group environments similar to the groups around \\rxj\\ that have velocity dispersions \\citep[$\\sim 400$~\\kmps,][]{tanaka2006} that are typical of groups found in the field \\citep{gerke2007}. However, the reversal in the SFR-density relation found by \\citet{elbaz2007} and \\citet{cooper2008} does not extend into the highest densities found in their group environments. Sample selection plays an important role in interpreting the different results. While we use a mass-limited sample, \\citet{elbaz2007} and \\citet{cooper2008} use luminosity-limited samples that are biased to include low-mass blue star-forming galaxies and exclude the corresponding non-star-forming red galaxies. When restricted to a mass-limited sample that was similar to the one in this work, \\citet{elbaz2007} found the SSFR continued to increase from low-density to high-density but with marginal significance (see Fig.~20 in that work). \\citet{cooper2008} found the mean SFR to increase with density by $\\sim 35\\%$ while for the same sample the mean SSFR {\\em decreases} by $\\sim 35\\%$ over the same density range, which implies an increase in the mean mass with density by a factor of $\\sim 2$. The inferred range of mean mass ($9.9 \\la \\log M/M_{\\odot} \\la 10.2$) from \\citet{cooper2008} is below our mass threshold. However, the small rise in SFRs with density seen by \\citet{cooper2008} can be explained by the presence of relatively more high mass galaxies, which have higher SFRs \\citep{elbaz2007,noeske2007b}, at higher densities in their sample. Interestingly, we note that when selecting a luminosity-limited sample that was similar to either of these works, we found the SFR to continue to decrease at higher densities, but at a much shallower pace for lower mass galaxies. This did not depend strongly on the sub-sample of galaxies that were used to compute local densities (i.e. luminosity vs. mass-limited). We note that contaminants from low-density environments, which we found to have higher levels of SF, likely contribute to the shallower trend. Recent observations of a ``star-forming sequence'' in the field at $z \\sim 0$ \\citep{salim2007} and up to redshifts of $z \\sim 1$ \\citep{noeske2007} suggest that galaxies populate a narrow distribution of SSFRs at a fixed stellar mass. Below this sequence is a more extended distribution of galaxies with very low SSFRs that are effectively non-star-forming. Here, we investigate whether the SSFR-density trend represents (1) a changing mix of galaxies that lie on or off of the star-forming sequence or (2) a change in the SFRs of star-forming galaxies in different environments. Without additional constraints, our MIPS stacking analysis alone cannot distinguish between these two scenarios. Using our previous results on the correlation between the red galaxy fraction and density \\citep{patel2009}, we can speculate on which of these scenarios is more likely. In \\citet{patel2009} we found the fraction of blue galaxies increased by a factor of $\\sim 10$ from $\\sim 5\\%$ in the highest density regions to $\\sim 50\\%$ at the lowest densities. If the typical SFR of galaxies on the star-forming sequence remains constant and the ratio of blue-to-red galaxies reflects the ratio of galaxies that are forming stars or not then one expects the average SFR (and SSFR within a single mass bin) in the high density regions to be diminished by a factor of $\\sim 10$ compared to the field, roughly what we see in the MIPS stacking analysis presented here. The notion that galaxies are either star-forming or not is given additional credence by \\citet{bell2005} and \\citet{dressler2009b} who found that the dominant mode of SF at these epochs was in moderate starbursts when accounting for the duty cycle of such events. In summary, we find that at $z \\sim 0.8$, the SSFRs and SFRs of galaxies with \\mcutrange\\ decrease in higher density environments, and this is true over the entire range of environments studied. This result follows the trend observed at $z \\sim 0$ in which \\citet{kauffmann2004} also found galaxies at higher densities to have lower SSFRs, but with SSFRs that were $\\sim 10$ times higher at $z \\sim 0.8$. Our result differs from that of \\citet{elbaz2007} and \\citet{cooper2008} who found a reversal in the SFR-density trend at $z \\sim 1$ over some portion of their density range, mostly driven by SF in lower mass galaxies present in their luminosity-limited samples. In a forthcoming paper, we plan to utilize rest-frame UV data with SED fitting to determine total SFRs of individual galaxies at $z \\sim 0.8$. The distribution of SFRs in different environments will allow us to probe the processes that were responsible for shutting down SF at a time when the universe was half its current age." }, "0910/0910.0006_arXiv.txt": { "abstract": " ", "introduction": "We observed a field centered on the asynchronous polar CD Ind (magnetic cataclysmic binary; see \\citeauthor{CDInd1}, \\citeyear{CDInd1} and \\citeauthor{CDInd2}, \\citeyear{CDInd2}) from December 21, 2008 until January 19, 2009 during 10 nights. Observations were done remotely at Tzec Maun Observatory (Pingelly, Western Australia) using the Takahashi TOA-150 apochromatic refractor (D=150 mm, F = 1095 mm) and SBIG STL-6303 CCD camera. All observations were conducted with the Bessel R filter. The field of view was $87' \\times 58'$. About 4 000 objects were detected on each frame. Our goal was to find new variable stars. ", "conclusions": "We examined light curves of 3712 objects in the $87' \\times 58'$ field centered on CD Ind and discovered three new variable stars. Two of them are most probably W UMa-type eclipsing binaries (EW), and the other one is most probably an RR Lyr-type (RRC) star. We show that \\emph{RMS-scatter vs. Magnitude} diagram is a useful tool to discover new variable stars. Table \\ref{table:summary} summarizes parameters of discovered variable stars, such as USNO-B1.0 designation, equatorial coordinates, brightness in maximum and minimum, period (in fractions of days), zero-epoch ($E_0$, time of maximum for RRC star and time of minimum for EW stars) and GCVS type of variability." }, "0910/0910.3763_arXiv.txt": { "abstract": "We derive kinematic distances to the 86 6.7 GHz methanol masers discovered in the Arecibo Methanol Maser Galactic Plane Survey. The systemic velocities of the sources were derived from \\coii~($J=2-1$), CS ($J=5-4$), and \\nhiii~observations made with the ARO Submillimeter Telescope, the APEX telescope, and the Effelsberg 100 m telescope, respectively. Kinematic distance ambiguities were resolved using \\hi~self-absorption with \\hi~data from the VLA Galactic Plane Survey. We observe roughly three times as many sources at the far distance compared to the near distance. The vertical distribution of the sources has a scale height of $\\sim 30$ pc, and is much lower than that of the Galactic thin disk. We use the distances derived in this work to determine the luminosity function of 6.7 GHz maser emission. The luminosity function has a peak at approximately $10^{-6}~L_\\sun$. Assuming that this luminosity function applies, the methanol maser population in the Large Magellanic Cloud and M33 is at least 4 and 14 times smaller, respectively, than in our Galaxy. ", "introduction": "Methanol masers at 6.7 GHz are unique compared to their OH and H$_2$O counterparts in that they appear to be exclusively associated with early phases of massive star formation. They are hence extremely useful tools to identify and study very young massive star forming regions. To date, over 800 sources have been detected through various targeted and blind surveys of the Galactic plane (e.g. \\citealt{gree09}; compilation of \\citealt{xu09b}). In spite of the extensive surveys to date for 6.7 GHz methanol masers, their luminosity function remains largely unknown. Since 6.7 GHz methanol masers are closely associated with mainline OH masers in massive star forming regions, \\citet{casw95} suggested that the luminosity function of methanol masers would be similar to that of OH masers, but with a scaling factor to account for the fact that methanol masers are on average much more intense than OH masers. \\citet{van96} use a probabilistic approach to assign distances and estimated the luminosity function to have a power-law behavior with an index around --2. Both these studies used only the peak flux density rather than the integrated flux for determining the luminosity functions. The most recent study of the issue was carried out by \\citet{pest07} who used a compilation of all sources detected prior to 2005 \\citep{pest05}. Carrying out a statistical analysis of the maser population and modeling the spatial distribution of the masers, the luminosity function was modeled by those authors as a power-law with sharp cutoffs and having an index between --1.5 and --2. There are two problems facing studies of the luminosity function. First, most studies use a catalog of methanol masers compiled from various surveys, many of them targeted (towards IRAS sources or OH masers), and with different sensitivities. However, unbiased searches have all shown that targeted searches, especially towards IRAS sources, underestimate the number of sources by a factor of 2 or greater \\citep{pand07b,szym02,elli96}. For the best estimate of the luminosity function, one should employ a blind survey, since it is possible to analyze the limitations such as completeness reliably. Among the several blind surveys to date, by far the most sensitive one is the Arecibo Methanol Maser Galactic Plane Survey (AMGPS; \\citealt{pand07a}). The AMGPS covered an area of 18.2 square degrees between Galactic longitudes of 35\\degr~and 54\\degr, and detected a total of 86 sources. This survey has a 95\\% probability of detection at a peak flux density of 0.27 Jy although sources as weak as 0.13 Jy have been detected. This makes the AMGPS catalog ideal for determining the luminosity function of 6.7 GHz methanol masers, especially at faint luminosities, although the relatively small size of the sample and area covered by the survey results in large statistical uncertainties. A more formidable problem is the determination of distances to the sources. Measuring distances is an old and challenging problem in observational astrophysics. While trigonometric parallax is the most reliable method for determining distances, it cannot be applied readily to a large sample of sources. The usual technique for Galactic sources is to use a Galactic rotation curve to determine kinematic distances. While kinematic distances can have significant errors at times (e.g. \\citealt{xu06}), recent work using Very Large Baseline Interferometry (VLBI) has discovered systematic proper motions in young massive star forming regions, which in principle can be taken into account to improve kinematic distance estimates \\citep{reid09}. A second challenge in the use of kinematic distances arises from an ambiguity between two distances (a ``near'' distance and a ``far'' distance) for sources located within the solar circle in the first and fourth Galactic quadrants. The most popular method to resolve the kinematic distance ambiguity is to measure an absorption spectrum towards the source (which has to have associated continuum radiation), and compare the velocities of the absorption lines with that of the source and the tangent point (see Fig. 1 of \\citealt{kolp03} for an illustration of this technique). Distance ambiguities have been successfully resolved towards ultracompact \\hii~regions using 21 cm \\hi~absorption \\citep{kuch94,kolp03,fish03} and 6 cm formaldehyde absorption \\citep{aray02,wats03,sewi04}. \\hi~absorption at 21 cm has been used by \\citet{pand08} to resolve the distance ambiguity towards 34 6.7 GHz methanol masers that have either directly associated 21 cm continuum, or belong to a cluster harboring a 21 cm continuum source. However, most methanol masers do not have any detectable radio continuum, presumably due to the young age of the exciting massive young stellar object. Hence, absorption line experiments can be used to resolve distance ambiguities towards only a small sample of 6.7 GHz methanol masers. \\citet{burt78} and \\citet{lisz81} discovered that several \\hi~self-absorption features in \\hi~maps of the Galactic plane correlated with CO emission features, and hence hypothesized that \\hi~self-absorption could be used to determine distances to molecular clouds. \\hi~self-absorption arises from cold \\hi~in the foreground absorbing warmer radiation of the background at the a specific radial velocity within the velocity range covered by the background gas. Hence, only molecular clouds at the near kinematic distance can display \\hi~self-absorption, since at the far distance there is no background emission at the radial velocity of the cloud. The theoretical work of \\citet{flyn04} shows that molecular clouds have enough opacity in cold \\hi~to exhibit self-absorption against strong 21 cm backgrounds (such as in the Galactic plane). \\hi~self-absorption has been used by \\citet{jack02} and \\citet{busf06} to resolve the distance ambiguity towards molecular clouds and massive young stellar candidates. In this paper, we describe work to resolve the distance ambiguity towards 6.7 GHz methanol masers detected in the AMGPS using \\hi~self-absorption. The sources were observed in CO ($J=2-1$) and NH$_3$ to determine their systemic velocities and the line profiles of thermally excited molecular emission. The maser emission of most sources have multiple emission components, which at times are spread over as much as 20~\\kms. While the central velocity of the maser emission is usually within a few \\kms~of the systemic velocity, it can at times be offset by more than 10 \\kms, and is hence not as reliable as velocities derived from thermal molecular emission. With distances determined from \\hi~self-absorption, we will then look at the distribution of sources in the Galaxy and derive the methanol maser luminosity function. This in turn will be used to compare the methanol maser population in our Galaxy with that in nearby external galaxies. ", "conclusions": "We used \\coii~($J=2-1$) observations and VGPS \\hi~data to resolve the kinematic distance ambiguity towards 6.7 GHz methanol masers discovered in the AMGPS. The distribution of the surface density of methanol masers as a function of Galactocentric distance agrees very well with existing estimates in the literature, although we observe the absolute numbers to be more than a factor of 3 higher. The vertical distribution of the sources has a scale height that is $\\sim 3$--5 times lower than that of the Galactic thin disk, presumably reflecting the smaller scale height of newly born massive stars. The resolution of the distance ambiguity allowed us to construct a reliable estimate of the luminosity function. Its shape does not agree with that of a power law, but has a peak around $\\sim 10^{-6}~L_\\sun$ followed by a decline towards lower luminosities. The luminosity function of 6.7~GHz methanol masers also appears to be different from that of mainline OH masers. Using the luminosity function, we derive estimates for the abundance of methanol masers in the LMC, SMC and M33 compared to the Milky Way. We find the under-abundance in M33 to be a factor of 3 higher than that in the LMC in spite of its higher metallicity. Finally, the distribution of sources between near and far distances closely follows the respective volumes sampled by the survey, thus indicating that the assumption of the near kinematic distance on a statistical basis should be avoided." }, "0910/0910.1766_arXiv.txt": { "abstract": "{} {The condensation of diffuse gas into molecular clouds and dense cores occurs at a rate driven largely by turbulent dissipation. This process still has to be caught in action and characterized.} {We observed a mosaic of 13 fields with the IRAM-PdB interferometer (PdBI) to search for small-scale structure in the \\twCO(1-0) line emission of the turbulent and translucent environment of a low-mass dense core in the Polaris Flare. The large size of the mosaic (1'$\\times$2') compared to the resolution (4'') is unprecedented in the study of the small-scale structure of diffuse molecular gas. } {The interferometer data uncover eight weak and elongated structures with thicknesses as small as $\\approx$ 3 mpc (600 AU) and lengths up to 70 mpc, close to the size of the mosaic. These are not filaments because once merged with short-spacings data, the PdBI-structures appear to be the sharp edges, in space and velocity-space, of larger-scale structures. Six out of eight form quasi-parallel pairs at different velocities and different position angles. This cannot be the result of chance alignment. The velocity-shears estimated for the three pairs include the highest values ever measured in regions that do not form stars (up to 780 \\kmspc). The CO column density of the PdBI-structures is in the range $N({\\rm CO)}=10^{14}$ to $10^{15}$\\cq\\ and their \\HH\\ density, estimated in several ways, does not exceed a few 10$^3$ \\cc. Because the larger scale structures have sharp edges (with little or no overlap for those that are pairs), they have to be thin layers of CO emission. We call them SEE(D)S for Sharp-Edged Extended (Double) Structures. These edges mark a transition, on the milliparsec scale, between a CO-rich component and a gas undetected in the \\twCO(1-0) line because of its low CO abundance, presumably the cold neutral medium. } {We propose that these SEE(D)S are the first directly-detected manifestations of the intermittency of interstellar turbulence. The large velocity-shears reveal an intense straining field, responsible for a local dissipation rate several orders of magnitude above average, possibly at the origin of the thin CO layers. } \\authorrunning{Falgarone et al.} \\offprints{E.~Falgarone} \\titlerunning{Extreme velocity-shears and CO on milliparsec scale} ", "introduction": "\\begin{figure} \\centering \\includegraphics[width=\\hsize{}]{10963f1.eps} \\caption{The location of the 13-field mosaic observed at the Plateau de Bure interferometer (centered at RA=01:55:12.26 and Dec=87:41:56.30) is shown as the box on top of the integrated emission of the \\twCO{} and \\thCO{} \\Jone{} maps obtained at the IRAM-30m. This is a place of low, almost featureless, CO line brightness. The arc-like structure visible in \\thCO\\ traces the outer layers of the low-mass dense core. Contour levels are shown in the wedges.} \\label{fig:location} \\end{figure} Turbulence in the interstellar medium (ISM) remains a puzzle in spite of dedicated efforts on observational and numerical grounds. This is because it is compressible, magnetized, and multi-phase, but also because of the huge range of scales separating those of injection and dissipation of energy. Moreover, because turbulence and magnetic fields are the main support of molecular clouds against their self-gravity, turbulent dissipation is a key process among all those eventually leading to star formation \\citep[see the reviews of][] {elmescalo04,scaloelme04}. In molecular clouds, turbulence is observed to be highly supersonic with respect to the cold gas. It is thus anticipated to dissipate in shocks in a cloud-crossing time (\\ie\\ $\\approx$ a few 10~Myr for giant molecular clouds of 100~pc with internal velocity dispersion of a few \\kms). Magnetic fields do not significantly slow the dissipation down \\citep{mmml98}. Actually, this is the basis of the turbulent models of star formation \\citep{maclow04} -- one of the two current scenarii of low-mass star formation -- in which self-gravitating entities form in the shock-compressed layers of supersonic turbulence. However, while it is unquestionable that the ISM is regularly swept by large-scale shock-waves triggered by supernovae explosions that partly feed the interstellar turbulent cascade \\citep{joung06,avillez07}, the smallest scales, barely subparsec in these simulations, are still orders of magnitude above the smallest observed structures and are unlikely to provide a proper description of the actual dissipation processes. Whether turbulent dissipation occurs primarily in compressive (curl-free) or in solenoidal (divergence-free) modes in the interstellar medium has therefore to be considered as an open issue. An ideal target to study turbulent dissipation is the diffuse molecular gas because it is the component in which dense cores form, with less turbulent energy density than their environment. The word ``diffuse'' here comprises all material in the neutral ISM at large that is not in dense cores \\ie\\ whose total hydrogen column density is less than a few $10^{21}$ \\cq. This includes the mixture of cold and warm neutral medium (CNM and WNM), the edges of molecular cloud complexes (also called translucent gas), and the high latitude clouds. Diffuse gas builds up a major mass fraction of the ISM. Actually, on the 30~pc scale, \\cite{goldsmith08} find that half the mass of the Taurus-Auriga-Perseus complex lies in regions having \\HH\\ column density below $2.1 \\times 10^{21}$ \\cq. Turbulent dissipation may also provide clues to the ``outstanding mysteries'' raised by observations of diffuse molecular gas \\citep[see the review of][]{snow06}: the ubiquitous small scale structure, down to AU-scales \\citep{heiles07}, the remarkable molecular richness found in this hostile medium, weakly shielded from UV radiation \\citep[e.g.][]{lilu98,gredel02}, the bright emission in the \\HH\\ pure rotational lines exceeding the predictions of photon-dominated region (PDR) models \\citep{falgarone05,lacour05}, the \\twCO\\ small-scale structures with a broad range of temperatures, \\HH\\ densities and linewidths that preclude a single interpretation in terms of cold dense clumps \\citep{ingalls2000, ingalls2007, heithausen04, heithausen06, sakamoto03}. The present paper extends the investigation of of turbulence down to the mpc-scale in the translucent environment of a low-mass dense core of the Polaris Flare. Over the years, this investigation has progressed along three complementary directions: \\\\ {\\it (i)} A two-point statistical analysis of the velocity field traced by the \\twCO\\ line emission, and conducted on maps of increasing size. Using numerical simulations of mildly compressible turbulence, \\cite{lis96} and~\\cite{pety03} first proposed that the non-Gaussian probability distribution functions ({\\it pdf}s) of line centroid velocity increments (CVI) be the signatures of the space-time intermittency of turbulence \\footnote{ Intermittency here refers to the empirical property of high Reynolds number turbulence to present an excess of rare events compared to Gaussian statistics, this excess being increasingly large as velocity fluctuations at smaller and smaller scales are considered \\citep[see the review of][]{anselmet01}. Although the origin of intermittency is still an open issue \\citep[but see][]{mordant02,chevillard05,arneodo08}, it is quantitatively characterized by the anomalous scaling of the high-order structure functions of the velocity and the shape of non-Gaussian {\\it pdf}s of quantities involving velocity derivatives \\citep[e.g.][]{frisch95}.} because the extrema of CVI (E-CVI) trace extrema of the line-of-sight average of the modulus of the {\\it pos} vorticity. Statistical analysis conducted on parsec-scale maps in two nearby molecular clouds have revealed that these extrema form parsec-scale coherent structures~\\citep[resp. Paper~III, HF09]{hily08,hf09}. \\\\ {\\it (ii)} A detailed analysis (density, temperature, molecular abundances) of these coherent structures, based on their molecular line emission. The gas there is more optically thin in the \\twCO\\ lines, warmer and more dilute than the bulk of the gas \\citep[hereafter Paper II]{hily07}, and large \\HCOp\\ abundances, unexpected in an environment weakly shielded from UV radiation, have been detected there \\citep[Paper~I]{falgarone06}.\\\\ {\\it (iii)} Chemical models of non-equilibrium warm chemistry triggered by bursts of turbulent dissipation~\\citep{joulain98}. The most recent progresses along those lines include the chemical models of turbulent dissipation regions (TDRs) by \\cite{godard09} and their successful comparison to several data sets, among which new submillimeter detections of \\thCHp(1-0) (Falgarone et al. in preparation). The \\twCO\\Jtwo\\ observations of the Polaris Flare with unprecedented angular resolution and dynamic range are the first to evidence the association between extrema of CVI and observed velocity-shears\\footnote{We use velocity-shear rather than velocity-gradient because the observations provide cross-derivatives of the velocity field, \\ie\\ the displacement measured in the plane-of-the-sky ({\\it pos}) is perpendicular to the line-of-sight velocity} (HF09). No shock signature (density and/or temperature enhancement, SiO detection) has been found in the coherent structure of E-CVI identified in the Polaris Flare (Hily-Blant and Falgarone, in preparation). All the above suggest (but does not prove yet) that the coherent structures carrying the statistical properties of intermittency are regions of intense velocity-shears where dissipation of turbulence is concentrated. The \\twCO(1-0) observations reported in this paper have been performed in a field located on one branch of the Polaris Flare E-CVI structure, in the translucent and featureless environment of a dense core (Fig.~\\ref{fig:location}). The outline of the paper is the following: the observations and data reduction are described in Section 2. The observational results are given in Section 3. The characterization of the emitting gas is made in Section 4 and we discuss, in Section 5, the possible origin and nature of the CO structures that we have discovered. Section 6 puts our results in the broad perspective provided by other data sets and Section 7 compares them to chemical model predictions and numerical simulations of turbulence. The conclusions are given in Section 8. ", "conclusions": "IRAM-PdBI observations of a mosaic of 13 fields in the turbulent environment of a low-mass dense core have disclosed small and weak \\twCO(1-0) structures in translucent molecular gas. They are straight and elongated structures but they are not filaments because, once merged with short-spacings data, the PdBI- structures appear as the sharp edges of larger-scale structures. Their thickness is as small as $\\approx$ 3 mpc (600 AU), and their length, up to 70 mpc, is only limited by the size of the mosaic. Their CO column density is a well determined quantity for the excitation conditions found at larger scale and is in the range $N({\\rm CO)}=10^{14}$ to $10^{15}$\\cq. Their \\HH\\ density, estimated in several ways, including the continuum emission of the brightest structure, does not exceed a few 10$^3$~\\cc. Their well-distributed orientations can be followed in the larger-scale environnement of the field. Six of them form three pairs of quasi-parallel structures, physically related. The velocity-shears estimated for the three pairs include the largest ever measured in non-star-forming clouds (up to 780 \\kmspc). The PdBI-structures are therefore not isolated and are the edges of so-called SEE(D)S for sharp-edged extended (double) structures. We show that the SEE(D)S are thin layers of CO-rich gas and that their sharp edges pinpoint a small-scale dynamical process, at the origin of the CO contrast detected by the PdBI. We propose that the SEE(D)S are the outcomes of the chemical enrichment driven by intense dissipation occurring in large velocity-shears and that they are CO-rich layers swept along by the straining field of CNM turbulence. The present work is the first detection of mpc-scale intense velocity-shears belonging to a parsec-scale shear. The large departure from average of the kinematic properties of these structures, confirms that they are a manifestation of the small-scale intermittency of turbulence in this high latitude field, a property already established on statistical grounds (HF09). The values of the velocity-shears (or rate-of-strain) provide a quantitative constraint on the dissipation rate that can be compared to chemical models. The link between the turbulent dissipation in the diffuse gas and the dense core observed in the vicinity of the PdBI mosaic (Fig.~\\ref{fig:location}) still remains to be established. Last, we would like to stress that sub-structure still exists in these mpc-scale structures of the diffuse ISM and that the next generation of interferometers (e.g. ALMA) should be able to observe gas at the dissipation scale of turbulence (that is still unknown) or at least observe the effects on the ISM (temperature, excitation, molecular abundances) of the huge release of energy expected to occur there." }, "0910/0910.1882_arXiv.txt": { "abstract": "We analyze the time dependence of fluid variables in general relativistic, magnetohydrodynamic simulations of accretion flows onto a black hole with dimensionless spin parameter $a/M=0.9$. We consider both the case where the angular momentum of the accretion material is aligned with the black hole spin axis (an untilted flow) and where it is misaligned by $15^\\circ$ (a tilted flow). In comparison to the untilted simulation, the tilted simulation exhibits a clear excess of inertial variability, that is, variability at frequencies below the local radial epicyclic frequency. We further study the radial structure of this inertial-like power by focusing on a radially extended band at $118(M/10M_\\odot)^{-1}{\\rm Hz}$ found in each of the three analyzed fluid variables. The three dimensional density structure at this frequency suggests that the power is a composite oscillation whose dominant components are an over dense clump corotating with the background flow, a low order inertial wave, and a low order inertial-acoustic wave. Our results provide preliminary confirmation of earlier suggestions that disk tilt can be an important excitation mechanism for inertial waves. ", "introduction": "Quasi-periodic oscillations (QPOs) are observed in the X-ray light curves of many black hole X-ray binaries (see \\citealt{rem06} for a recent review). They have also been observed in extragalactic ultraluminous X-ray sources \\citep{str03,str07} and in one active galactic nucleus \\citep{gie08}. The origin of these phenomena is still far from clear. One class of models centers on trapped wave modes within the accretion flow. In particular, a variety of modes have been proposed in hydrodynamic models of geometrically thin accretion disks (e.g. \\citealt{wag99,kat01}). Axisymmetric inertial-acoustic modes (``$f$-'' or ``inner $p$-modes'') and axisymmetric inertial modes (``$g$-'' or ``$r$-modes'') \\footnote{Throughout this paper we will use the term inertial mode and $r$-mode interchangeably. We prefer to avoid the use of the term $g$-mode, whose primary restoring force is generally due to entropy gradients. In disks, the primary restoring force for inertial modes is due to specific angular momentum gradients.} can be trapped due to the existence of the innermost stable circular orbit around a black hole, which produces a maximum in the radial profile of the radial epicyclic frequency. Non-axisymmetric inertial modes can also be radially trapped between inner and outer Lindblad resonances. The very existence of inertial modes in accretion disks has been challenged recently, as numerical simulations appear to show that magnetorotational (MRI) turbulence suppresses them. This has been demonstrated in local shearing box simulations \\citep{arr06}, whose boundary conditions would artificially trap a discrete axisymmetric mode spectrum. Power spectra from these simulations show discrete acoustic modes and a radial epicyclic oscillation, but no inertial modes. This behavior has also been seen in global simulations of accretion disks in a pseudo-Newtonian potential: hydrodynamic disks exhibit trapped axisymmetric inertial modes, whereas MHD turbulent disks do not \\citep{rey08}. It appears that axisymmetric inertial modes, which necessarily have frequencies at or below the local radial epicyclic frequency, are particularly vulnerable to nonlinear damping by MRI turbulence, which has a power spectrum that peaks near the orbital frequency. Non-axisymmetric inertial modes can have higher frequencies and might be less vulnerable to MRI turbulence, but there has as yet been no convincing demonstration of the existence of a trapped inertial mode spectrum in any MHD simulation. Another reason why discrete inertial modes may not be present in accretion disks with MRI turbulence is that even subthermal magnetic fields can extend the inner trapping radius of the propagation zone of both axisymmetric and non-axisymmetric inertial modes down to the innermost stable circular orbit (ISCO) \\citep{fu09}. Unless inertial waves can reflect off the plunging region of the flow, standing waves will no longer be sustainable. On the other hand, it has been suggested that warps and eccentricity in disks may play a fundamental role in a nonlinear excitation mechanism of inertial modes in geometrically thin accretion disks \\citep{kat04a,kat08,fer08}. These large scale deformations interact with trapped $r$-modes giving rise to intermediate modes. These intermediate modes then couple back to the warp or eccentricity to produce positive feedback on the original $r$-mode oscillations (see \\citealt{fer08} for a more detailed explanation of this coupling mechanism). \\cite{fra07} recently completed fully general relativistic, MHD simulations of accretion disks with misaligned black hole spin and disk angular momentum vectors (``tilted disks''). These tilted disks are strongly warped near the black hole and exhibit global epicyclic oscillations superimposed on the turbulence in the flow \\citep{fra08}. These oscillations manifest themselves as eccentric orbits of fluid elements in the disk, albeit with a $180^\\circ$ flip in orientation of the elliptical orbits across the midplane of the disk. Tilted disks may therefore provide favorable conditions for the nonlinear excitation of inertial modes. Here we report the results of a search for trapped modes in simulations of both an untilted and tilted disk. In agreement with previous work, we do not observe the presence of modes in the untilted simulation. However, at the same numerical resolution, we observe significant excess power with frequencies characteristic of inertial waves in the tilted disk. Because of the complexity of the background flow, the physical nature of this power is difficult to determine. Nevertheless, its spatial structure appears to confirm that this is partly inertial in character. This may represent preliminary confirmation of the warp/eccentricity excitation mechanism, showing that it can be strong enough to overcome damping due to MRI turbulence. It also shows that trapping, in the sense of radial localization of power, can be maintained even in the presence of magnetic fields. This paper is organized in the following manner. In section 2 we briefly review the simulation parameters and discuss our power spectrum analysis of the simulation data. Section 3 presents power spectra from both untilted and tilted geometries, evidence that the latter may be dominated by inertial-like variability, and an analysis of the three dimensional structure of the power at a particular frequency within this inertial regime. We summarize our conclusions in section 4. ", "conclusions": "Previous work has demonstrated that inertial modes are vulnerable to disruption by MRI turbulence \\citep{arr06,rey08}. Our Fourier analysis of the untilted simulation, a configuration devoid of tilt or eccentricity, corroborates this conclusion. However, nonlinear analytic calculations \\citep{kat04a, kat08, fer08} have indicated that disk warps and eccentricity may excite inertial modes; both of these features are generically present in tilted accretion flows. Though the analysis of the variability in our tilted simulation at $118{\\rm Hz}$ does not appear to be wholly inertial (or even acoustic) in nature, the presence of two weak, odd parity, $m=0$ and $m=2$ components provides some support for this hypothesis. Evidence that disk tilt is connected with the excitation of inertial or acoustic waves is twofold. First, fundamentally different shapes in the spectra from the untilted and tilted simulations implies a correlation between tilt and the power at frequencies characteristic of inertial waves. Second, the radial structure of the variability suggests a superposition of modes, at least one of which may be described as inertial in nature in the context of relevant analytic models. Together, these clues establish the presence of variability unique to a tilted geometry which appears to be at least partially inertial in character. Recalling that a tilted disk configuration naturally yields a low frequency QPO activity in black hole X-ray binaries and speculating that our work's inertial-like variability may be related to high frequency QPO activity in the same systems, we advocate further numerical exploration of these tilted geometries. Torques associated with the Kerr spacetime can produce a bodily precession of the tilted disk at frequencies within the $0.1$ and $30{\\rm Hz}$ range associated with the low frequency QPO. Because this precession frequency is strongly dependent on the radial extent of the disk in this model, it readily explains the observed correlation between the LFQPO's frequency and disk flux \\citep{rem06,fra08,ing09}. Similarly, our current work shows inertial or acoustic variability excited and maintained by the disk tilt that exhibits frequencies of order the radial epicyclic frequency maximum. Such frequencies fall within the observed range of $40$ to $450{\\rm Hz}$ for HFQPOs in stellar mass black hole systems. HFQPOs are only observed when a system's X-ray emission exhibits the characteristics of the steep power-law state (cf. \\citealt{rem06} for a review of the properties of X-ray black hole binaries), and we must acknowledge, however, that our simulations are almost certainly not accurate representations of that state. For instance, since they neither account for radiative cooling nor capture all forms of dissipation, our simulations likely do not reproduce some of the key features found in various models for the state's structure (e.g. the energetically significant corona in \\citealt{don07}). However, we are confident that our simulations and analysis provide insight into the dynamical effects of a tilted accretion flow, particularly, the excitation of inertial and acoustic modes. \\sloppypar{ We thank Chris Done, Paul Henisey, Mami Machida, Phil Marshall, Gordon Ogilvie, Chris Reynolds, and Alexander Tchekhovskoy for useful conversations, and Sam Cook for his help with data transport and preparation. We also thank the anonymous referee for comments that significantly improved this paper. This work was supported in part by NSF grant AST-0707624. The work of BTF was supported by FCT (Portugal) through grant SFRH/BD/22251/2005. PCF acknowledges support from NSF grant AST-0807385 and an American Astronomical Society International Travel Grant. We are also grateful to the Nordic Institute for Theoretical Astrophysics for hosting a workshop on QPOs where much of this work was completed. }" }, "0910/0910.1599_arXiv.txt": { "abstract": "The joint likelihood of observable cluster signals reflects the astrophysical evolution of the coupled baryonic and dark matter components in massive halos, and its knowledge will enhance cosmological parameter constraints in the coming era of large, multi-wavelength cluster surveys. We present a computational study of intrinsic covariance in cluster properties using halo populations derived from Millennium Gas Simulations (MGS). The MGS are re-simulations of the original $500 \\hinv \\mpc$ Millennium Simulation performed with gas dynamics under two different physical treatments: shock heating driven by gravity only ($\\GO$) and a second treatment with cooling and preheating ($\\PH$). We examine relationships among structural properties and observable X-ray and Sunyaev-Zel'dovich (SZ) signals for samples of thousands of halos with $M_{200} \\ge 5 \\times 10^{13} \\hinv\\msun$ and $z < 2$. While the X-ray scaling behavior of \\PH model halos at low-redshift offers a good match to local clusters, the model exhibits non-standard features testable with larger surveys, including weakly running slopes in hot gas observable--mass relations and $\\sim 10\\%$ departures from self-similar redshift evolution for $10^{14} \\hinv\\msun$ halos at redshift $z \\sim 1$. We find that the form of the joint likelihood of signal pairs is generally well-described by a multivariate, log-normal distribution, especially in the \\PH case which exhibits less halo substructure than the \\GO model. At fixed mass and epoch, joint deviations of signal pairs display mainly positive correlations, especially the thermal SZ effect paired with either hot gas fraction ($r=0.88/0.69$ for $\\PH/\\GO$ at $z=0$) or X-ray temperature ($r=0.62/0.83$). The levels of variance in X-ray luminosity, temperature and gas mass fraction are sensitive to the physical treatment, but offsetting shifts in the latter two measures maintain a fixed $12\\%$ scatter in the integrated SZ signal under both gas treatments. We discuss halo mass selection by signal pairs, and find a minimum mass scatter of $4\\%$ in the \\PH model by combining thermal SZ and gas fraction measurements. ", "introduction": "Accurate cosmology using surveys of clusters of galaxies requires a robust description of the relations between observed cluster signals and underlying halo mass. Even without strong prior knowledge of the mass-signal relation, cluster counts, in combination with other probes, add useful constraining power to cosmological parameters \\citep{cunha:09a}. However, significant improvements can be realized when the error in mass variance is known \\citep{limaHu:05, cunha:09b}. Improvements can also be gained by extending the model to multiple observed signals \\citep{cunha:08}, especially when an underlying physical model can effectively reduce the dimensionality of the parameter sub-space associated with the model \\citep{younger:06}. The coming era of multiple observable signals from combined surveys in optical, sub-mm and X-ray wavebands invites a more holistic approach to modeling multi-wavelength signatures of clusters. Signal covariance characterizes survey selection, in terms of mass and additional observables. For the case of X-ray selected samples, \\cite{nord:08} demonstrate that luminosity--temperature covariance can mimic apparent evolution in the luminosity--mass relation under analysis that combines deep, X-ray-flux limited samples with local, shallow ones. Employing a selection observable with small mass variance minimizes such errors. Recent work has shown that the total gas thermal energy, $Y$, observable via an integrated Sunyaev-Zeldovich (SZ) effect \\citep{carlstrom:04} or via X-ray imaging and spectroscopy, is a signal that scales as a power-law in mass with only $\\sim 15\\%$ scatter \\citep{white:02, kravtsov:06, maughan:07, ohara:07, zhang:08, jeltema:08}. However, unbiased estimates of the mass selection function for $Y$ or any other signal requires accurate knowledge of how the signal--mass scaling relation evolves with redshift. The redshift behavior of signals is generally not well known empirically, although recent work has begun to probe evolution in X-ray signals to $z \\sim 1$ \\citep{maughan:06,vikhlinin:08}. Emerging samples from wide-area SZ surveys should dramatically improve this situation. One can address signal--mass covariance using hydrostatic, virial or lensing mass estimates from observations, but several sources of systematic and statistical error challenge this approach. For hydrostatic masses, early gas dynamic simulations \\citep{evrard:90,nfw, thomas:97} suggested that turbulent gas motions drove hydrostatic masses to underestimate true values by $\\sim 20\\%$. More sophisticated recent models, with a factor thousand improvement in mass resolution, demonstrate this effect at a similar level in the mean, with $\\sim 15\\%$ scatter among individual systems \\citep{rasia:06, nagai:06, jeltema:08}. Cluster masses can also be measured by the shear induced on background galaxies due to gravitational lensing. With this method, individual cluster masses have mass uncertainties $\\sim 20\\%$ due to cosmic web confusion \\citep{hoekstra:03, deputter:05}, but large samples can reduce the uncertainty in the mean. The mean scaling behavior of samples binned in some selection signal offers another empirical path to measuring covariance. Non-zero covariance between the selection and an independent, follow-up signal implies that the selection-binned scaling relation of the followup signal with mass need not match that signal's intrinsic mass scaling. Comparison of scaling relations from differently selected samples thereby offers insight into covariance. \\cite{rykoff:08} offer a first attempt at this exercise for X-ray luminosity and optical richness using the optically-selected SDSS \\maxbcg\\ sample \\citep{koester:07}. The sample contains $\\sim 13,000$ clusters for which weak lensing mass estimates have been made by stacking the shear of richness-binned sub-samples \\citep{sheldon:07,johnston:07}. \\cite{rykoff:08} stack Rosat All-Sky Survey data \\citep{voges:99} in the same \\maxbcg\\ richness bins, and find that the mean X-ray luminosity--mass relation derived with richness binning is consistent at the $\\sim 2 \\sigma$ level with relations derived solely from X--ray data \\citep{reiprich:02, stanek:06}. A theoretical approach to studying cluster covariance is to realize populations via numerical simulation. While high resolution treatment of astrophysical processes, including star formation, supernova and AGN feedback, galactic winds and thermal conduction have been included in recent simulations \\citep{dolag:05, borgani:06, kravtsov:06, sijacki:08, puchwein:08}, the computational expense has limited sample sizes to typically a few dozen objects. A detailed study of population covariance requires larger sample sizes, as can be generated by lower resolution simulations of large cosmological volumes using a more limited physics treatment \\citep{bryan:98, hallman:06, gottlober:07}. We take the latter approach in this paper, focusing on the bulk properties of massive halos identified the Millennium Gas Simulations (MGS), a pair of resimulations of the original $500 \\hinv$ Mpc Millennium run \\citep{springel:05}, each with $10^9$ total particles, half representing gas and half dark matter. The pair of runs use different treatments for the gas physics --- a gravity-only ({\\texttt{GO}}) simulation, sometimes called ``adiabatic'', in which entropy is increased via shocks, and a simulation with cooling and preheating ({\\texttt{PH}}). The former ignores galaxies as both a sink for baryons and a source of feedback for the hot intracluster medium (ICM). The latter also ignores the mass fraction contribution of galaxies, but it approximates the feedback effects of galaxy formation by a single parameter, an entropy level imposed as a floor at high redshift \\citep{evrard:91, kaiser:91, bialek:01, kay:07, gottlober:07}. Our study focuses on samples of $\\sim 5000$ halos with mass $M > 5 \\times 10^{13} \\hinv \\msun$ examined at multiple epochs covering the redshift range $0 < z < 2$. The paper is organized as follows: in Section \\ref{sec:sims} we discuss the details of the Millennium Gas Simulations and our halo finding approach. In Section \\ref{sec:scaling}, we present mean scaling relation behavior in both mass and redshift. We then turn to covariance about the mean scaling relations in Section \\ref{sec:covar}. Unless otherwise noted, our units of mass are $10^{14} \\hinv \\msun$ and halo mass is defined within a sphere encompassing a density contrast $\\Delta_c = 200$ times the critical density. ", "conclusions": "\\label{sec:con} We analyze scaling relations for multiple properties of massive halos taken from a pair of gas dynamic simulations with different gas physics treatments. Our samples contain tens of thousands of halos with masses $M_{200} > 5 \\times 10^{13} \\hinv \\msol$ at redshifts $z \\lta 2$. The physical treatments of gravity only ($\\GO$) or preheating ($\\PH$) are both highly idealized, but we show that the latter reproduces the scaling relation behavior of core-extracted X-ray measures of local clusters. The dark matter velocity dispersion scales with mass and redshift according to self-similar expectations, indicating that the virial theorem is respected regardless of gas treatment. However, the gas behavior in both treatments differs from self-similarity. The deviations in the \\GO case tend to be small and are related to the mass-limited sample definition. The entropy injection of the \\PH model drives more substantial deviations from self-similar scaling. At $z=0$, the ICM mass fraction varies with mass in a manner roughly consistent with observed measurements of local clusters. We find that $\\ficm$ requires a quadratic fit in log-mass, and mild curvature in the logarithmic scalings of $Y$ and $\\elbol$ as a function of mass is also evident. While the ICM in \\PH halos is lower in mass compared to the \\GO case, it is also slightly hotter. The effects on $\\ficm$ and $T_m$ nearly cancel when combined to form the thermal SZ signal, leading to nearly identical $Y$--$M$ scaling relations above $3 \\times 10^{14}$ for both \\PH and \\GO cases at low redshift. The ICM mass fraction at fixed mass declines weakly with redshift, by $10\\%$ in $5 \\times 10^{14} \\hinv\\msol$ halos at $z=1$ and with larger reductions at lower masses. While {\\sl Chandra\\/} observations of optically-selected clusters in the RCS survey show evidence of reduced gas fractions at $z=1$ \\citep{hicks:06}, further work is needed to address the quantitive level of agreement between the observations and \\PH model expectations. The \\PH baryon fraction evolution drives departures from self-similar predictions in the $Y$--$M$ and $\\elbol$--$M$ relations; slopes tend to steepen slightly and the normalization at $10^{14} \\hinv\\msol$ is lower than self-similar expectations at high redshift. We present the first systematic investigation of property covariance in the computational samples of massive halos. The data generally support a multivariate log-normal form for the joint distribution of signals at fixed mass and redshift. All measures depart somewhat from an exact gaussian form, but the deviations in gas measures are smaller for the \\PH model due to the suppression of substructure caused by the preheating. Most signal pairs exhibit positive correlations, with the lone exception of $-0.1$ between $\\sigmadm$ and $\\ficm$ in the \\PH case. The thermal SZ signal displays a robust $13\\%$ scatter that is strongly correlated with variations in both ICM gas mass and temperature, with $\\ficm$ dominating in the \\PH case and $T_m$ being more important in the \\GO treatment. Combining multiwavelength observations offers an opportunity to improve selection of clusters by their intrinsic mass. We derive the mass variance of signal pairs and show that combining strongly correlated signals always improves mass selection, even when one of the signals by itself is a comparatively poor mass proxy. The combination of thermal SZ and ICM mass fraction in the \\PH case selects halo mass with just $4\\%$ intrinsic rms scatter. Identifying the root causes behind the terms in the covariance matrix is a considerable task that we leave for future work. Mergers \\citep{roettigerBurnsLoken:97, rickerSarazin:01} and assembly bias (\\cite{Boylan-Kolchin:09} and references therein) will surely play important roles for many cluster signals. Studies of merger history behavior will shed light on survey selection properties, particularly potential biases related to the dynamical state of a halo. The simple treatment of baryon physics in our simulations limits our investigation to the hot, thermal ICM of clusters. More extensive physical treatments that incorporate galaxy and supermassive black hole formation and other physics such as MHD and non-thermal plasmas will ultimately extend the set of observable halo signals into the optical/NIR and radio wavebands." }, "0910/0910.1550_arXiv.txt": { "abstract": "In the weak field limit general relativity reduces, as is well known, to the Newtonian gravitation. Alternative theories of gravity, however, do not necessarily reduce to Newtonian gravitation; some of them, for example, reduce to Yukawa-like potentials instead of the Newtonian potential. Since the Newtonian gravitation is largely used to model with success the structures of the universe, such as for example galaxies and clusters of galaxies, a way to probe and constrain alternative theories, in the weak field limit, is to apply them to model the structures of the universe. In the present study we consider how to probe Yukawa-like potentials using $N$-body numerical simulations. \\PACS{04.50.+h, 04.50.Kd, 45.50.-j} ", "introduction": "\\label{intro} Einstein's General Relativity (GR) is one of the most beautiful theories ever imagined by the human mind. Although GR is a successful theory of gravitation, it is unable to explain, for example, the accelerating expansion of the universe, unless a cosmological constant or a dark energy fluid is considered. Nowadays other theories of gravitation intend to give an alternative interpretation to this accelerating expansion (see, e.g., \\cite{piazzamarinoni}). Most of these alternative theories, although having different approaches (some scalar-tensor theories of gravitation, nonsymmetric gravitational theory, etc), reduce, in the weak-field limit, to a Yukawa-like gravitational potential (hereafter YGP), i.e., \\begin{equation} \\phi(r)=-\\frac{GM}{r}e^{-r/{\\lambda}}. \\label{yukpot} \\end{equation} The above equation gives the potential of a point mass m at a distance r; G is the universal gravitational constant and $\\lambda$ is a constant. When $\\lambda \\rightarrow \\infty$, this potential tends to the Newtonian one. Note that the parameter $\\lambda$ is the Compton wavelength of the exchange particle, which in the present case is a graviton. The graviton mass is related to $\\lambda$ through the well known equation $m_g=h / \\lambda c$, where $h$ is the Planck constant and c is the speed of light. The YGP has been investigated in the literature, in particular, in galactic astronomy and cosmology. For example, Signore \\cite{signore} studies this potential under the cosmological context, on the other hand, de Araujo \\& Miranda \\cite{araujoemiranda2007} study how variations of the $\\lambda$ parameter can disturb galactic disks. This last study generalizes the (Newtonian) Toomre's \\cite{toomre} expressions for rotation curves in order to account for a YGP. Rodriguez-Meza et al. \\cite{rodriguezmeza} consider the YGP as a correction on the Newtonian potential. Recently, some numerical approaches have been developed to study alternative theories. Cervantes-Cota et al. \\cite{cervantes2007a,cervantes2007b} have developed numerical studies of a scalar-tensor theory in the weak field limit, using the $N$-body method. In these studies they consider that the gravitational potential in the weak field limit is given by $$\\Phi_{STT} = \\phi_N(r) + \\alpha \\phi(r)/ (1+ \\alpha)\\, , $$ \\par\\noindent where $\\Phi_{STT}$ is the potential from the scalar-tensor theory, $\\phi_N(r)$ is the Newtonian potential, $\\phi(r)$ is given by Eq.(1), and $\\alpha \\equiv 1/(3 + 2 \\omega)$, with $\\omega$ being the Brans-Dicke parameter \\cite{brans}. They changed a particular $N$-body code replacing the Newtonian potential by the potential $\\Phi_{STT}$ and analyzed its influence on isolated galaxies, two colliding spiral galaxies and issues concerning the formation of bars. Simulations of cosmological structure formation in $\\rm{\\Lambda}$CDM scenarios are also studied. In the present paper, we use similar techniques to that used by Cervantes et al. \\cite{cervantes2007a,cervantes2007b}, although with a gravitational potential derived from other theories. We modify a popular $N$-body code, \\textbf{Gadget-2} \\cite{springel2005}, \\textsl{replacing} the Newtonian potential by a pure YGP as given by Equation \\ref{yukpot}, which is derived from some alternative theories of gravitation (see, e.g., Visser's theory \\cite{visser}). It is worth stressing that although we have chosen a particular gravitational potential, the approach considered here can be applied to any alternative gravitational potential. This paper is organized as follows: in section 2, we show the changes made in the \\textbf{Gadget-2} code, in section 3, we test the YGP \\textbf{Gadget-2} code using a pair of particles and $N$-body systems. Finally, in section 4, we present the conclusions and briefly discuss how this modified code can be used to test alternative theories of gravitation via galactic dynamics. ", "conclusions": "The present paper considers the use of the well known \\textbf{Gadget-2}, with the appropriate modifications, to study alternative theories of gravitation, with particular attention to the YGP. In this first paper we discuss how to modify the \\textbf{Gadget-2} and evaluate how reliable this modified code is, performing a series of tests. As we have shown in the previous sections, our tests are successful: numerical errors from our changes on \\textbf{Gadget-2} are neglected and we can use this modified code to probe alternative theories of gravitation. For example, we can use arguments based on galactic dynamics to exclude or not the Yukawa potential hypothesis. This is very important, due to the fact that alternative theories appear frequently in the literature, but so far there are no many investigations in the galactic scales using $N$-body codes and galactic dynamics formalism. We will show, in future papers, how to test these alternative theories using early and late-type galaxies." }, "0910/0910.3887_arXiv.txt": { "abstract": "We investigate whether a tachyonic scalar field, encompassing both dark energy and dark matter-like features will drive our universe towards a Big Brake singularity or a de Sitter expansion. In doing this it is crucial to establish the parameter domain of the model, which is compatible with type Ia supernovae data. We find the 1$\\sigma$ contours and evolve the tachyonic sytem into the future. We conclude, that both future evolutions are allowed by observations, Big Brake becoming increasingly likely with the increase of the positive model parameter $k$. ", "introduction": "With the discovery of cosmic acceleration \\cite{cosm} the quest for modeling dark energy \\cite{dark} has started. Besides the most simple cosmological constant, other models based on various perfect fluids with negative pressure, like Chaplygin gas \\cite{Chaplygin}, minimally and non-minimally coupled scalar fields and fields having non-standard kinetic terms \\cite% {kinetic,tachyons} were advanced. The latter ones include as a subclass the models based on different forms of the Born-Infeld-type action, which is often associated with the tachyons arising in the context of string theory \\cite{string}. Due to the non-linearity of the dependence of the tachyon Lagrangians on the kinetic term of the tachyon field, the dynamics of the corresponding cosmological models appears to be very rich. The tachyon model studied in paper \\cite{we-tach} contains a 2-fluid analogue scalar field $T$, the dynamics of which is given by a simple potential, depending on two parameters, $\\Lambda $ and $k$. The model is homogeneous and isotropic. A phase space diagram in the tachyonic field and its derivative $s\\equiv \\dot{T}$ shows 5 type of distinct cosmological evolutions possibly occurring for the model, some of them containing regimes where $s$ is superluminal. All evolutions originate from one of the Big Bangs of the model, but they either end in a de Sitter infinite exponential expansion, as the $\\Lambda $CDM model does, or in a future singularity characterized by a regular scale factor $a$, vanishing Hubble parameter $H$ and energy density $\\varepsilon $, but infinite $s$ and pressure $p$. Most notably, the second time derivative of the scale factor goes to $-\\infty $, the reason why we call this singularity a Big Brake. A kinematical analysis \\cite{Barrow} predicted the existence of such singularities, named sudden future singularities. From a combined kinematical and observational reasoning alone, sudden future singularities could occur as early as in ten million years \\cite{Dabrowski}, however no underlying dynamics is known to support this. Classically the Big Brake singularity is stable. This can be seen by a series expansion of the scale factor in the vicinity of the singularity and checking the stability conditions advanced in Ref. \\cite{Barrow-priv}. Its quantum study indicated singularity avoidance \\cite{KKS}. Recently \\cite{tachyon-prd} the compatibility of the model with type Ia supernovae observation has been investigated. After we present some basic features of the model in Section 2, in Section 3 we give more details on this compatibility check, in terms of the original variables employed in Ref. \\cite{we-tach}. Then in Section 4 we stress the crucial difference between negative and positive values of the model parameter $k$. While for the former all evolutions end in the de Sitter attractor, for positive $k$ the 1$\\sigma $ contour compatible with type Ia supernovae contains both states which evolve into de Sitter or into a Big Brake. In this dynamical model the Big Brake can occur no earlier than $10^{8}$ years. The Big Brake singularity belongs to the class of soft cosmological singularities, which also includes other representants \\cite{soft}. Other types of singularities arising in the study of various dark energy models include the Big Rip singularity \\cite{Rip}, present in some models with phantom dark energy \\cite{phantom}. The possibility of existence of a phase of contraction of the universe, ending up in the standard Big Crunch cosmological singularity was also considered \\cite{Crunch}. \\textit{Unit convention: }the Newtonian constant is normalized as $8\\pi G/3=1 $ and we take $c=1$. ", "conclusions": "" }, "0910/0910.4233_arXiv.txt": { "abstract": "We revisit the statistical significance of the ``dark flow'' presented in \\cite{kashlinsky09}. We do not find a statistically significant detection of a bulk flow. Instead we find that CMB correlations between the 8 WMAP channels used in this analysis decrease the inferred significance of the detection to 0.7$\\sigma$. ", "introduction": "\\label{sec:intro} A recent set of papers \\citep{kashlinsky08, kashlinsky09} claims to have detected the velocities of galaxy clusters with respect to the cosmic microwave background (CMB) frame by means of the kinetic Sunyaev-Zel'dovich (kSZ) effect \\citep{sunyaev72}. The papers suggest the existence of a ``dark flow'': a 700 km s$^{-1}$ bulk flow of all matter out to a redshift of at least $z\\simeq 0.1$ ($\\rm{r}\\simeq400$ Mpc). The magnitude and direction of the flow are claimed to be consistent with the peculiar velocity of the Local Group with respect to the CMB frame as inferred from the CMB dipole \\citep{kogut93}. Velocity coherence over such large scales is not predicted by the standard $\\Lambda\\rm{CDM}$ cosmology and would, if confirmed, constitute a major observational result. In this paper we revisit the analysis presented in \\cite{kashlinsky09}, hereafter referred to as K09. The K09 analysis seeks to measure the kSZ signal of a sample of $\\sim$700 X-ray-selected galaxy clusters. The 3-year WMAP temperature maps for 8 differencing assemblies \\citep{hinshaw07} are high-pass filtered in an attempt to remove the primary CMB anisotropy. The temperatures of the filtered maps at the galaxy cluster locations are fit to a dipole, which is interpreted as the kSZ signal induced by a bulk flow of the galaxy clusters. We will argue that the uncertainty of this measurement is dominated by primary CMB anisotropy, not detector noise. As the CMB is observed by all 8 WMAP channels, the errors are highly correlated between these channels, and the inferred detection significance is greatly reduced. ", "conclusions": "\\label{sec:conclusion} We have revisited the analysis presented in \\cite{kashlinsky08, kashlinsky09} which reports a significant detection of a bulk flow of $\\sim$700 galaxy clusters out to $z\\simeq0.1$ by means of the kSZ effect. We have demonstrated that the estimates for the kSZ signal are highly correlated between the different WMAP channels used in this analysis and that this correlation is caused by primary CMB anisotropy. We have simulated the errors on the kSZ measurement while taking into account these CMB correlations and find that there is not a significant detection of a kSZ signal or bulk flow. \\begin{figure*} \\begin{center} \\includegraphics[width=1.0\\textwidth]{err.pdf} \\caption{1000 simulated estimates for the 3 dipole components. These simulations take into account the CMB correlations between the different WMAP channels. The uncertainty is highest on the dipole components that lie in the galactic plane ($a_x$ and $a_y$) because of the geometry of the galactic mask.} \\label{fig:err} \\end{center} \\end{figure*} \\begin{figure*} \\begin{center} \\includegraphics[width=0.7\\textwidth]{corr.pdf} \\caption{1000 simulated estimates for the $a_x$ dipole component from two randomly chosen WMAP channels, Q1 and W2. The estimates are highly correlated ($\\rho=0.9$). This high level of correlation is common to all pairs of channels and is caused by primary CMB fluctuations.} \\label{fig:corr} \\end{center} \\end{figure*}" }, "0910/0910.3139_arXiv.txt": { "abstract": "A few star clusters in the Magellanic Clouds exhibit composite structures in the red-clump region of their colour--magnitude diagrams. The most striking case is NGC~419 in the Small Magellanic Cloud (SMC), where the red clump is composed of a main blob as well as a distinct secondary feature. This structure is demonstrated to be real and corresponds to the simultaneous presence of stars which passed through electron degeneracy after central-hydrogen exhaustion and those that did not. This rare occurrence in a single cluster allows us to set stringent constraints on its age and on the efficiency of convective-core overshooting during main-sequence evolution. We present a more detailed analysis of NGC~419, together with a first look at other populous Large Magellanic Cloud clusters which are apparently in the same phase: NGC~1751, NGC~1783, NGC~1806, NGC~1846, NGC~1852 and NGC~1917. We also compare these Magellanic Cloud cases with their Galactic counterparts, NGC~752 and NGC~7789. We emphasise the extraordinary potential of these clusters as {\\em absolute} calibration marks on the age scale of stellar populations. ", "introduction": "\\begin{figure}[b] \\begin{center} \\includegraphics[width=0.8\\textwidth]{figure_lr.eps} \\caption{False-colour images (in the electronic version) of the Small Magellanic Cloud star cluster NGC~419, derived from ACS/WFC {\\it (bottom left)} and HRC {\\it (top right)} images in the F555W and F814W filters. The top left and bottom right panels zoom in onto the HRC image. Red-clump stars are marked with circles. At first glance, it is evident that they are quite homogeneous in their colours and luminosities, as expected for red-clump stars. However, when comparing their first Airy rings, one notices some quite subtle and {\\em systematic} differences in their brightnesses. Indeed, there are two kinds of red-clump stars: the most numerous and brighter, and a subsample (about 15\\% of the total) of fainter `secondary red-clump stars' (marked with red and green circles, respectively, in the electronic version). The luminosity difference between these groups is about 0.4~mag. There is also a $\\sim0.04$~mag difference in the mean colour which, however, is too small to be noticeable in the figure. The two kinds of red-clump stars are uniformly distributed across the HRC image. } \\label{fig_image} \\end{center} \\end{figure} NGC~419 is a populous star cluster located to the east of the Small Magellanic Cloud (SMC)'s bar in a region relatively devoid of dust and free from contamination by the SMC field. There are two wonderful pairs of {\\sl HST} images of this cluster, taken in the F555W and F815W filters, originally obtained as part of programme GO-10396 (PI J. S. Gallagher) and now retrievable from the {\\sl HST} archive. They are shown in Figure~\\ref{fig_image}. While the ACS/WFC images reveal the overall cluster structure, the ACS/HRC observations show the details of the cluster core. \\begin{figure}[b] \\begin{center} \\includegraphics[width=0.6\\textwidth]{cmd_hrc.ps} \\caption{NGC~419 colour--magnitude diagram (CMD; {\\it left panel}), derived from the ACS/HRC data (Girardi et al. 2009). The error bars in the left panel are upper limits to the photometric errors. The two right-hand panels detail the red-clump and main-sequence turnoff, clearly showing their composite structure.} \\label{fig_cmd} \\end{center} \\end{figure} The HRC images allow us to perform photometry of an impressive quality, and reveal at least two surprises: the presence of multiple main sequence turnoffs (MMSTOs), and a composite red clump which contains a pronounced faint extension (Figure~\\ref{fig_cmd}). Both features were noticed by Glatt et al. (2008). They tentativeley interpreted the faint red-clump structure as being caused by SMC field contamination. However, this explanation does not stand up to a simple star count in the neighbouring field selected from the ACS/WFC image (see Girardi et al. 2009): the expected number of field red-clump stars in the ACS/HRC area is just 4.5, while the faint red clump contains at least 50 objects (Figure~\\ref{fig_cmd}). Assuming a Poissonian distribution for the field stars, the probability ($P$) that these 50 stars are drawn from the SMC field alone is virtually zero ($P<10^{-9}$). Detached binaries cannot give rise to this faint red clump either, since any combination of single stars will cause an extension of the red clump to brighter magnitudes. What, then, is the dual red clump of NGC~419 made of? Girardi et al. (2009) show that it is caused by the presence of stars following two very different evolutionary paths. In the following, we will expand their arguments a little. There are, in fact, a few different ways of looking at dual red clumps, as detailed below. ", "conclusions": "" }, "0910/0910.5145_arXiv.txt": { "abstract": "We have derived the quantum vacuum pressure $p_{\\rm vac}$ as a primary entity, removing a trivial and a gauge terms from the cosmological constant-like part (the zeroth term) of the effective action for a matter field. The quantum vacuum energy density $\\tilde{\\varrho}_{\\rm vac}$ appears a secondary entity, but both are of expected order. Moreover $p_{\\rm vac}$ and $\\tilde{\\varrho}_{\\rm vac}$ are dynamical, and therefore they can be used in the Einstein equations. In particular, they could dynamically support the holographic dark energy model as well as the ``thermodynamic'' one. ", "introduction": " ", "conclusions": "" }, "0910/0910.2848_arXiv.txt": { "abstract": "We investigate the relationship between the linewidths of broad \\mgii\\ $\\lambda$2800 and \\hb\\ in active galactic nuclei (AGNs) to refine them as tools to estimate black hole (BH) masses. We perform a detailed spectral analysis of a large sample of AGNs at intermediate redshifts selected from the Sloan Digital Sky Survey, along with a smaller sample of archival ultraviolet spectra for nearby sources monitored with reverberation mapping (RM). Careful attention is devoted to accurate spectral decomposition, especially in the treatment of narrow-line blending and \\feii\\ contamination. We show that, contrary to popular belief, the velocity width of \\mgii\\ tends to be smaller than that of \\hb, suggesting that the two species are not cospatial in the broad-line region. Using these findings and recently updated BH mass measurements from RM, we present a new calibration of the empirical prescriptions for estimating virial BH masses for AGNs using the broad \\mgii\\ and \\hb\\ lines. We show that the BH masses derived from our new formalisms show subtle but important differences compared to some of the mass estimators currently used in the literature. ", "introduction": "It is generally accepted that active galactic nuclei (AGNs) are powered by the release of gravitational energy from material accreted onto supermassive black holes (BHs). The determination of BH mass (\\mbh) is crucial for understanding the AGN phenomena, the cosmological evolution of BHs, and even the coevolution of AGNs and their host galaxies. Yet, for such distant objects, it is currently impossible to obtain direct measurement of \\mbh\\ using spatially resolved stellar or gas kinematics. Fortunately, significant advances have been made in recent years from reverberation mapping (RM) studies of nearby Seyfert galaxies and quasi-stellar objects (QSOs; e.g., Wandel et al. 1999; Kaspi et al. 2000; Peterson et al. 2004). First, the anti-correlation between the radius of the broad-line region (BLR) and the velocity width of broad emission lines for single objects supports the idea that the BLR gas is virialized and that its velocity field is dominated by the gravity of the BH (Peterson \\& Wandel 1999, 2000; Onken \\& Peterson 2002). Second, the size of the BLR scales with the continuum luminosity (Kaspi et al. 2000, 2005), approximately as $R \\propto L^{0.5}$ (Bentz et al. 2006, 2009); the $R-L$ relation offers a highly efficient procedure for estimating the BLR size without carrying out time-consuming RM observations. And third, the BH masses estimated by RM are roughly consistent (Gebhardt et al. 2000b; Ferrarese et al. 2001; Nelson et al. 2004; Onken et al. 2004) with the predictions from the tight correlation between \\mbh\\ and bulge stellar velocity dispersion established for inactive galaxies (the \\mbh--$\\sigma_\\star$ relation; Gebhardt et al. 2000a; Ferrarese \\& Merritt 2000). These developments imply that we can estimate the BH mass in type 1 (broad-line, unobscured) AGNs by simple application of the virial theorem, \\mbh\\ = $fRv^2/G$, where $f$ is a geometric factor of order unity that depends on the geometry and kinematics of the line-emitting region, $R$ is the radius of the BLR derived from the AGN luminosity, and $v$ is some measure of the virial velocity of the gas measured from single-epoch spectra. The feasibility of obtaining $R$ and $v$ from single-epoch spectra enables \\mbh\\ to be estimated very efficiently for large samples of AGNs, especially for luminous quasars at higher redshift that typically exhibit only slow and small-amplitude variability (e.g., Kaspi et al. 2007), with the assumption that the virial relation is independent of redshift and can be extrapolated to higher luminosities and masses. In practice, for those AGNs that have measurements of $\\sigma_\\star$, $f$ is determined empirically by scaling the virial masses to the \\mbh--$\\sigma_\\star$ relation of inactive galaxies (e.g., Onken et al. 2004). Implicit in this practice is the assumption---one open to debate (Greene \\& Ho 2006; Ho et al. 2008; Kim et al. 2008)---that active and inactive BHs should follow the same \\mbh--$\\sigma_\\star$ relation. The most widely used estimator for $v$ is the full width at half-maximum (FWHM) of the line. Now, a large number of formalisms to estimate \\mbh\\ from single-epoch spectra have been proposed in the recent literature, using different broad emission lines optimized for different redshift regimes probed by (widely available) optical spectroscopy. At low redshifts, the lines of choice are \\hb\\ (Kaspi et al. 2000; Collin et al. 2006; Vestergaard \\& Peterson 2006) or \\ha\\ (Greene \\& Ho 2005). At intermediate redshifts, \\mgii\\ $\\lambda$2800 is used (McLure \\& Jarvis 2002), while at high redshifts, one has to resort to \\civ\\ $\\lambda$1549 (Vestergaard 2002; Vestergaard \\& Peterson 2006). These formalisms are ultimately calibrated against RM masses based on the \\hb\\ BLR radius and linewidth measured from the variable (rms) spectra (Peterson et al. 2004). Because the \\hb\\ linewidth is typically smaller in the rms spectra than in the single-epoch or mean spectra (Vestergaard 2002; Collin et al. 2006; Sulentic et al. 2006), some authors have proposed that the FWHM used in the \\hb-based formalisms should be further corrected to obtain unbiased \\mbh\\ estimates (Collin et al. 2006; Sulentic et al. 2006). As for the \\mgii-based formalisms, because there are very few RM experiments of the \\mgii\\ line, they are either based on the RM data for \\hb\\ (e.g., McLure \\& Jarvis 2002; McLure \\& Dunlop 2004) or calibrated against the \\hb\\ formalisms themselves (e.g., Kollmeier et al. 2006; Salviander et al. 2007). A strong, underlying assumption is that \\mgii\\ and \\hb\\ are emitted from the same location in the BLR and have the same linewidth (see also Onken \\& Kollmeier 2008). In support of this assumption, some authors find that \\mgii\\ and \\hb\\ indeed have very similar linewidths (e.g., McLure \\& Jarvis 2002; McLure \\& Dunlop 2004; Shen et al. 2008; also cf. Salviander et al. 2007). However, there are conflicting results in the literature: Corbett et al. (2003) claimed that \\mgii\\ is generally broader than \\hb, whereas Dietrich \\& Hamann (2004) came to an opposite conclusion. Certainly, the most direct way to settle this issue is through direct RM of the \\mgii\\ line. So far there are only two objects that have successful \\mgii\\ RM, NGC\\,5548 (Clavel et al. 1991; Dietrich \\& Kollatschny 1995) and NGC\\,4151 (Metzroth et al. 2006). These studies tentatively suggest that \\mgii\\ responds more slowly to continuum variations than \\hb, implying that the \\mgii-emitting region is larger than that radiating \\hb. Thus, there are still some important open questions regarding the robustness of \\mbh\\ measurements based on \\mgii. What is the relation between the linewidths of \\hb\\ and \\mgii? Are estimates of \\mbh\\ based on \\mgii\\ consistent with those based on \\hb? These basic questions are critical for understanding the systematic uncertainties in studies of the cosmological evolution of BHs (cf. Shen et al. 2008; McGill et al. 2008; Denney et al. 2009a). To address the above questions, we perform a detailed comparison of the widths of the \\mgii\\ and \\hb\\ lines using single-epoch spectra for a large, homogeneous sample of Seyfert 1 nuclei and QSOs at intermediate redshifts culled from the Sloan Digital Sky Survey (SDSS; York et al. 2000). We further compare single-epoch \\mgii\\ linewidths with \\hb\\ linewidths measured from the rms spectra of AGNs with RM observations, finding systematic deviations between the two. We present a recalibration of the \\mgii\\ virial mass estimator and compare our formalism with previous ones in the literature. This paper adopts the following set of cosmological parameters: $H_{\\rm 0}$=70 km\\,s$^{-1}$\\,Mpc$^{-1}$, $\\Omega_m $=0.3, and $\\Omega_{\\rm \\Lambda}$=0.7. ", "conclusions": "We investigate the relation between the velocity widths for the broad \\mgii\\ and \\hb\\ emission lines, derived from FWHM measurements of single-epoch spectra from a homogeneous sample of 495 SDSS Seyfert 1s and QSOs at $0.45 < z <0.75$. Careful attention is devoted to accurate spectral decomposition, especially in the treatment of narrow-line blending and \\feii\\ contamination. We find that \\mgii\\ FWHM is systematic smaller than \\hb\\ FWHM, such that FWHM(\\mgii)\\,$\\propto$\\,FWHM(\\hb)$^{0.81\\pm 0.02}$. Using 29 AGNs that have optical RM data and usable archival UV spectra, we then investigate the relation between single-epoch \\mgii\\ FWHM and rms \\hb\\ \\sigline\\ (line dispersion), a quantity regarded as a good tracer of the virial velocity of the BLR clouds emitting the variable \\hb\\ component. We find that, similar to the situation for the FWHM of single-epoch \\hb, single-epoch \\mgii\\ FWHM is unlikely to be linearly proportional to rms \\hb\\ \\sigline. The above two findings suggest that a major assumption of previous \\mgii-based virial BH mass formalisms---that the \\mgii-emitting region is identical to that of $\\hb$---is problematic. This finding and the recent updates of the reverberation-mapped BH masses (Peterson et al. 2004; Denney et al. 2006, 2009b; Metzroth et al. 2006; Bentz et al. 2007; Grier et al. 2008) motivated us to recalibrate the \\mbh\\ estimator based on single-epoch \\mgii\\ spectra. Starting with the empirically well-motivated BLR radius--luminosity relation and the virial theorem, $\\mbh \\propto L^\\beta {\\rm FWHM}^\\gamma$, we fit the reverberation-mapped objects in a variety of different ways to constrain $\\beta$ and $\\gamma$. For all the strategies we have considered, $\\beta$ has a well-defined value of $\\sim$0.5, in excellent agreement with the latest BLR radius--luminosity relation (Bentz et al. 2006, 2009), whereas $\\gamma \\approx 1.5 \\pm 0.5$, which is marginally in conflict with the canonical value of $\\gamma = 2$ normally assumed in past studies. Performing a similar exercise for \\hb\\ yields $\\mbh \\propto L^{0.5} {\\rm FWHM(\\hb)}^{1.09\\pm0.22}$, which again significantly departs from the functional forms used in the literature. The 1 $\\sigma$ uncertainty (scatter) is of the order of 0.3 dex relative to the RM-based masses for the \\hb\\ estimator, and $\\sim$0.4 dex for the \\mgii\\ estimator. Using the same data set, the scatter of our \\hb\\ mass scaling relation is reduced by 0.1 dex over that of Vestergaard \\& Peterson (2006), indicating improvement in the internal scatter. We use the SDSS database to compare our new \\mbh\\ estimators with various existing formalisms based on single-epoch \\hb\\ and \\mgii\\ spectra. BH masses derived from our \\mgii-based mass estimator show subtle but important deviations from many of the commonly used \\mbh\\ estimators in the literature. Most of the differences stem from the recent recalibration of the masses derived from RM. Researchers should exercise caution in selecting the most up-to-date \\mbh\\ estimators, which are presented here." }, "0910/0910.2765_arXiv.txt": { "abstract": "We find statistically significant spatial correlations between the arrival directions of the highest energy cosmic rays (HECRs) observed by the Akeno Giant Air Shower Array (AGASA) and large-scale structure of galaxies observed by Sloan Digital Sky Survey (SDSS) in the redshift ranges of $0.006 \\leq z < 0.012$ and $0.012 \\leq z < 0.018$ at angular scale within $\\sim 5^{\\circ}$. This result supports a hypothesis that the sources of HECRs are related to galaxy distribution even in the northern sky, which has been already indicated by Pierre Auger Observatory in the southern sky. We also investigate the dependency of the correlation on the absolute magnitude, color, and morphology of the galaxies. For galaxies with $0.006 \\leq z < 0.012$, the correlation tends to be stronger for luminous and red galaxies. Based on these results, we discuss plausible HECR sources and constraint on Galactic magnetic field. ", "introduction": "\\label{introduction} Although the sources of the highest energy cosmic rays (HECRs) have been poorly known, recent observations begin to unveil a part of the nature of HECR origin. The first indication to HECR sources was reported by the Akeno Giant Air Shower Array (AGASA). The AGASA reported the anisotropic distribution of cosmic rays (CRs) above $4 \\times 10^{19}$ eV at small angular scale comparable with their angular uncertainty to determine arrival directions of primary CRs \\cite{takeda99}. The anisotropic feature has been interpreted as the existence of point-like extragalactic sources. An important step to their origin is the positional correlation between the arrival directions of HECRs with energies above $\\sim 6 \\times 10^{19}$ eV and nearby astrophysical objects reported by the Pierre Auger Observatory (PAO) \\cite{abraham07,abraham08}. These results strongly suggested that HECRs should be of extragalactic origin. Several groups have also analyzed the PAO data by using catalogs of different astrophysical objects and then have confirmed the correlation with matter (or galaxy) distribution of local Universe \\cite{kashti08,george08,ghisellini08,takami09c}. Although the same analysis of new PAO data as Refs. \\cite{abraham07,abraham08} shows the decreases of the significance level of the correlation \\cite{hague09}, the correlation with the large-scale structure is still positive \\cite{aublin09}. The PAO observes HECRs from the southern sky, since it is located in Argentina. If HECR sources are extragalactic objects as suggested by the PAO, the correlation of HECRs and galaxy distribution is also expected in the northern sky. However, such correlation has never been clearly reported, though the correlation of HECRs with the supergalactic plane was reported in the northern sky \\cite{stanev95,uchihori00}. We tested the correlation of the AGASA data with galaxies of Infrared Astronomical Satellite Point Source Redshift Survey (IRAS PSCz) \\cite{saunders00}, but a significantly positive correlation could not be found \\cite{takami09c}. There are two reasons why we have not found such correlation when we assume that the correlation exists in the northern sky. One is the small exposure of the AGASA relative to the PAO. In the period of Refs. \\cite{abraham07,abraham08}, the total exposure of the PAO is $\\sim$ 7 times as large as that of the AGASA. Thus, the AGASA may not have the number of highest energy (HE) events large enough to find correlation. The other is the completeness of a galaxy catalog used in \\cite{takami09c}. In other words, it is the possibility that the IRAS PSCz catalog of galaxies does not reflect the large-scale structure of galaxy distribution in local Universe. In this study, we focus the latter reason, because the former problem cannot be solved by us without accumulating more number of HECR events. We investigate the correlation between HECR events observed in the northern sky and galaxy distribution which has higher completeness than the IRAS PSCz catalog. For this purpose, we adopt galaxies observed by Sloan Digital Sky Survey (SDSS) \\cite{york00}. The SDSS is a survey project of galaxies to investigate the large-scale structure of the Universe in optical bands and therefore can observe faint galaxies down to $m_r \\sim 20.0$ where $m_r$ is the $r$-band apparent magnitude of galaxies. Thus, SDSS galaxies better trace the structure of the local Universe. As HECR events observed in the northern sky, we adopt the AGASA data published in Ref. \\cite{hayashida00}. This paper is laid out as follows: in Section \\ref{methods}, we explain the data samples of HECR events and galaxies in detail. Also, a statistical method which we adopt to test the correlation between them is described. In Section \\ref{results}, we calculate cross-correlation functions defined in Section \\ref{methods} between the AGASA events and SDSS galaxies with several redshift ranges. We also investigate the dependencies of the correlation on the $r$-band absolute magnitude, color, and morphology of galaxies. These give us useful information to search for the HECR sources because the features of the host galaxies reflect the nature of objects contained in the hosts. Section \\ref{discussion} is devoted to discuss and interpret the results obtained in Section \\ref{results}. Finally, we conclude this study in Section \\ref{conclusion}. ", "conclusions": "\\label{conclusion} In this study, we investigated the correlation between the arrival directions of HECRs detected in the northern sky (the AGASA data) and galaxy distribution observed by the SDSS. The SDSS can observe faint galaxies, and therefore SDSS galaxies are a good sample to investigate the correlation with the large-scale structure of matter in local Universe. We found the positive correlation of the AGASA events with SDSS galaxies within the angular scale of $\\sim 5^{\\circ}$ with the redshift ranges of $0.006 \\leq z < 0.012$ and $0.012 \\leq z < 0.018$ for the first time. This result supported a hypothesis that the sources of HECRs are related to galaxy distribution even in the northern sky, which has been already indicated by several analyses of the PAO data in the southern sky \\cite{abraham07,abraham08,kashti08,george08,ghisellini08,takami09c}. We also tested the dependencies of the correlation signals on the r-band absolute magnitude, color, and morphology of the galaxies, since these information is generally useful to constrain HECR sources involved in the host galaxies. For galaxies with $0.006 \\leq z < 0.012$, we found that the AGASA events correlated with luminous ($M_r \\leq -19$) and red ($M_g - M_r > 0.6$) galaxies, and the luminous galaxies around the AGASA events were relatively red and morphologically late-type. Such dependencies were not clear for galaxies with $0.012 \\leq z < 0.018$. This was possible because the flux of HECRs from a sources with this redshift range is smaller than for $0.006 \\leq z < 0.012$ and/or the trajectories of HECRs are more deflected than for sources with $0.006 \\leq z < 0.012$. We discussed possible source candidates based on our results for galaxies with $0.006 \\leq z < 0.012$ in Section \\ref{discussion}. Unfortunately, the interpretation of the results is ambiguous at present because of uncertainty of relations between HECR source candidates and the nature of their host galaxies, and our poor knowledge on intervening magnetic fields. Under the limitations, possible interpretation is summarized as follows: Morphology dependence of galaxies correlating with the AGASA events is positive for GRBs and magnetars, but the colors of these galaxies are relatively red, contrary to relatively blue colors expected for these objects. Note that the discussion on the colors was only qualitative. More detail and quantitative discussion is required to confirm/reject GRB or magnetar origin of HECRs. On the other hand, HECR generation by non-thermal phenomena related to AGNs might be difficult from results in this study because the morphologies of luminous galaxies near the AGASA events are relatively late-type, though such phonomena are strongly related to galaxies with early-type galaxies. The interpretation above is based on the assumption that the trajectories of HECRs are not largely deflected by IGMF. When IGMF is strong, fake correlation could occur \\cite{kotera08,ryu09}. The correlation between the arrival directions of HECRs and the positions of their sources is reduced, though the correlation with matter distribution is conserved. In this case, it is not possible to confirm HECR sources by the nature of galaxies correlating with HECR events. The angular scale of the positive correlation enabled us to constrain intervening magnetic fields under the assumption that we see the correlation between HE protons and their sources. The angular scale estimated in this study, $< 5^{\\circ}$, was compared with theoretical predictions by Ref. \\cite{takami09f}. The angular scale was consistent with predictions based on bisymmetric spiral field models of GMF. Since the larger angular deflections of HE protons are predicted for axisymmetric GMF models, it might be difficult to realize the axisymmertic models. In order to investigate correlation signals between HECRs and matter distribution/their sources more accurately, we need more number of HECR events. It was not clear whether the correlation features found in this study for galaxies with $0.006 \\leq z < 0.012$ are common in the Universe. This can be checked by increasing the observed number of HECRs. We can also understand intervening magnetic fields and the nature of HECR sources. Ref. \\cite{takami08a} predicted the spatial correlation between HE protons and their sources in local Universe even taking a realistically structured IGMF model into account, which was clearly visible by accumulating $\\sim 200$ events above $\\sim 6 \\times 10^{19}$ eV for the number density of HECR sources of $10^{-5}$ Mpc$^{-3}$. Data accumulation in the northern sky is in progress. The total exposure of the Telescope Array have already reached 75\\% of the AGASA exposure \\cite{taketa09} and is expected to reach that of the AGASA in this year. Thus, their data is quite useful to test the correlation and the nature of the host galaxies of HECR sources. Also, two projected HECR observatories with extremely large effective area, Extreme Universe Space Telescope (JEM-EUSO) \\cite{ebisuzaki09} and the northern site of the PAO \\cite{harton09}, enable us not only to clarify the correlation but also to unveil the positions of HECR sources in the northern sky of local Universe. \\ack We are grateful to K.Maeda and I.Kayo for useful discussion. This work is supported by Grants-in-Aid for Scientific Research from the Ministry of Education, Culture, Sports, Science and Technology (MEXT) of Japan through No.21840019 (H.T.) and No.19104006 (K.S.). The work of T.N. is supported by the JSPS fellow. This work is also supported by World Premier International Research Center Initiative (WPI Initiative), from the MEXT of Japan. Funding for the SDSS and SDSS-II has been provided by the Alfred P. Sloan Foundation, the Participating Institutions, the National Science Foundation, the U.S. Department of Energy, the National Aeronautics and Space Administration, the Japanese Monbukagakusho, the Max Planck Society, and the Higher Education Funding Council for England. The SDSS Web Site is http://www.sdss.org/. The SDSS is managed by the Astrophysical Research Consortium for the Participating Institutions. The Participating Institutions are the American Museum of Natural History, Astrophysical Institute Potsdam, University of Basel, University of Cambridge, Case Western Reserve University, University of Chicago, Drexel University, Fermilab, the Institute for Advanced Study, the Japan Participation Group, Johns Hopkins University, the Joint Institute for Nuclear Astrophysics, the Kavli Institute for Particle Astrophysics and Cosmology, the Korean Scientist Group, the Chinese Academy of Sciences (LAMOST), Los Alamos National Laboratory, the Max-Planck-Institute for Astronomy (MPIA), the Max-Planck-Institute for Astrophysics (MPA), New Mexico State University, Ohio State University, University of Pittsburgh, University of Portsmouth, Princeton University, the United States Naval Observatory, and the University of Washington. \\clearpage" }, "0910/0910.5659_arXiv.txt": { "abstract": "Hamiltonian perturbation theory is used to analyse the stability of $f(R)$ models. The Hamiltonian equations for the metric and its momentum conjugate are written for $f(R)$ Lagrangian in the presence of perfect fluid matter. The perturbations examined are perpendicular to $R$. As perturbations are added to the metric and momentum conjugate to the induced metric instabilities are found, depending on the form of $f(R)$. Thus the examination of these instabilities is a way to rule out certain $f(R)$ models. ", "introduction": "The question of dark energy has been at the heart of cosmology since the discovery of accelerating expansion of the universe \\cite{riess98}. The traditional picture of general relativity with ordinary relativistic or non-relativistic matter in homogeneous and isotropic universe meets severe problems when accommodating it to current cosmological observations. The conflicting observational evidence comes mainly from supernova light curves \\cite{riess98,perlmutter99}, CMB anisotropies \\cite{Bennett2003,Netterfield2001} and large scale structures \\cite{Perlmutter1999,Bunn1996}. This has lead to several suggested remedies. Perhaps the most popular way is to add some non-conventional matter to the universe. Among these the simplest possibility is no doubt to use the cosmological constant. A review of the subject can be found in \\cite{peebles2002}. In any case, the key aspect is the negative pressure of the new matter which boosts the expansion of the universe. Other considerations include more general distribution of matter, {\\it i.e.} non-homogeneous or non-isotropic universe (see {\\it e.g.} \\cite{Alnes2005}). Besides these two, a lot of effort has been put into studies on generalizations and modifications of General Relativity. For example metric-affine theories (see {\\it e.g.} \\cite{Sotiriou2006}), scalar-tensor theory (see {\\it e.g.} \\cite{faraoni04,Bransdicke1961}), brane-world gravity (see {\\it e.g.} \\cite{Maartens2003}) and more general Lagrangians have been considered. In the present paper we are especially interested in $f(R)$ gravity models in which the Einstein-Hilbert action is replaced by a function of the curvature scalar $R$ \\cite{Carroll2003,Carroll2004,Meng2003, Allemandi2004,Sotiriou2008}. None of these modifications is free of problems and this is indeed the case of $f(R)$ gravity as well. As for any model, the cosmological observations issue some constraints (see {\\it e.g.} \\cite{Tsujikawa2007,Starobinsky2007}) as do the observations in the solar system (see {\\it e.g.} \\cite{Chiba2003, Hu2007,Magnano93,multamaki2008,Henttunen2007}). The opinions are still divided on the viability of $f(R)$ theories of gravity. There are numerous approving studies (see {\\it e.g.} \\cite{Faraoni2006,Olmo2006}) as well as sceptical ones (see {\\it e.g.} \\cite{Faulkner2006,Erickcek2006}). As the actual universe is not homogeneous and isotropic but contains local perturbations, additional challenges for $f(R)$ theories emerge from stability analysis \\cite{Dolgov2003,Soussa2003,Faraoni2005}. An acceptable cosmological model has to be stable against perturbations in the metric and the mass distribution. However, stability analysis is customarily done only in the direction of $R$, {\\it i.e.} only curvature perturbations are considered. This is motivated in particular in the case of General Relativity, where the relation between space-time curvature and the matter density is a simple one: the trace of Einstein equations imply $R\\propto \\rho$. This in turn implicates a simple and direct relation between the perturbations in matter and curvature. This is not the only possibility. In a $f(R)$ model the model the relation is more complicated due to appearance of function $f(R)$ and higher derivative terms in the field equations. The phase space is considerably larger and metrics corresponding a given matter distribution ambiguous. The physical acceptability, however, of a model requires general stability; also stability against perturbations which keep curvature constant, perpendicular to $R$. The Hamiltonian formulation of general relativity has been around since the work of Arnowitt, Deser and Misner \\cite{Arnowitt1962}\\footnote{The ideas were first seen in the long out of print {\\it Gravitation: an introduction to current research}. The authors have later on released the article on ArXiv as cited.}. Hamiltonian formulation has also surfaced in the works of Ashtekar \\cite{ashtekar87}. The first papers on the subject often neglected the boundary terms, however, later works have clarified these details ({\\it e.g.} \\cite{Brown1992, Hawking1995}). Hamiltonian formulation has not received too much interest in contemporary papers. In particular and to our knowledge the use of Hamiltonian formulation on perturbations of $f(R)$ theories has not been studied so far. The main interest has been in specific choices for the function $f(R)$. In the present paper we look into perturbations using Hamiltonian formalism of $f(R)$ theories. While the technique has not yet been applied to general $f(R)$ theories with perturbations it is a useful tool in studying the stability of $f(R)$ models: with it is simple to study perturbations perpendicular to $R$. As in classical mechanics the Hamiltonian is written as a functional of the fields and their canonical momenta. However, in a geometric theory like General Relativity and $f(R)$ theories, some complications appear due to constraints between field components. The two main aspects of the canonical Hamiltonian formalism are that the field equations are of the first order in the time derivatives and that time is distinguished from other coordinates. For writing the Hamiltonian equations we must thus foliate the region of space-time with space-like hypersurfaces. Finally the resulting field equations for the perturbations are then analyzed for instabilities. The conventions and details of the formalism can be found in \\cite{poisson2004}. The paper is organized as follows. In section \\ref{hamform} we write the Hamiltonian field equations. The 3+1 decomposition and foliation of the space-time are also presented. The first order perturbations are added to the system in section \\ref{1order}. We also take a look at second order perturbations in section \\ref{2order}. In section \\ref{concl} we summarize our results. ", "conclusions": "\\label{concl} Traditionally the stability analysis is performed in the Lagrangian formalism and the analysis parallel to $R$ has been carried out before in several papers ({\\it e.g.} \\cite{Dolgov2003}). Many of the interesting $f(R)$ models have been found to be inherently unstable in the past \\cite{Kainulainen20071,Faraoni2005}. However, stability analysis has not yet been used to the full extent as long as the studies concentrate on curvature perturbations only. By using the Hamiltonian instead of Lagrangian formulation we examined the large scale cosmological perturbations perpendicular to $R$ with non-relativistic matter. These perturbations are fairly easy to examine with the Hamiltonian formulation. The found instabilities are noticeably different to those of previous works ({\\it e.g.} \\cite{Faraoni2005}). Because of the constraint $\\dot R=0$ diagonal perturbations of metric and conjugate momentum are related to each other. The temporal part of the metric showed up to be constrained by the conjugate momentum one. Moreover, the spatial part of the metric is forced to vanish. If these constraints were not satisfied we would have non-trivial perturbations of non-diagonal elements of the metrics like $g_{0i}$. The perturbations of the momentum conjugate turn out to be the most interesting ones. The equation depends explicitly on the form of the function $f(R)$. Some choices of $f(R)$ lead clearly unstable cosmological model, but seemingly not all. We have studied some well-known $f(R)$ functions and find them unstable. Albeit the physical interpretation of the perturbation momentum conjugate is unfortunately not as clear as that of the metric perturbations, equation \\eqref{pab} demonstrates the relation between the momentum conjugate and the extrinsic curvature. In the 3+1 decomposition the intrinsic curvature $\\tilde R$ defines how the hypersurface is curved whereas the extrinsic curvature defines how each slice is curved relative to the enveloping space-time. As the perturbations were not well-behaved in this context further studies would be relevant in order to find the limits of these constraints. Fruitful directions would likely to be investigating the effects of more other types of matter. Also, it would be prudent to examine the case where the metric can include shift ({\\it i.e.} $N_a\\neq 0$). It is clear from the form of \\eqref{hamden} that such a generalization would affect the following equations of motion deeply as the last term would be non-zero. This is understandable as the metric would now include spatio-temporal elements. It is also possible to study more general theories with the Lagrangian depending also on for example $R_{\\mu\\nu}R^{\\mu\\nu}$ or Gauss-Bonnet term. It appears that with Hamiltonian formulation of perturbations can be used to constrain the spectrum of cosmologically acceptable $f(R)$ theories. While there are several physical arguments to judge the $f(R)$ theories like cosmological observations and solar system behaviour, stability analysis is one important tool to rule out ill-behaved models out of numerous possible modified theories of gravity. With continued studies it is possible to find the ones best describing the observed behaviour of the universe. \\subsection*" }, "0910/0910.0774_arXiv.txt": { "abstract": "Compared to starburst galaxies, normal star forming galaxies have been shown to display a much larger dispersion of the dust attenuation at fixed reddening through studies of the IRX-$\\beta$ diagram (the IR/UV ratio ``IRX'' versus the UV color ``$\\beta$''). To investigate the causes of this larger dispersion and attempt to isolate second parameters, we have used GALEX UV, ground-based optical, and Spitzer infrared imaging of 8 nearby galaxies, and examined the properties of individual UV and 24~$\\mu$m selected star forming regions. We concentrated on star-forming regions, in order to isolate simpler star formation histories than those that characterize whole galaxies. We find that 1) the dispersion is not correlated with the mean age of the stellar populations, 2) a range of dust geometries and dust extinction curves are the most likely causes for the observed dispersion in the IRX-beta diagram 3) together with some potential dilution of the most recent star-forming population by older unrelated bursts, at least in the case of star-forming regions within galaxies, 4) we also recover some general characteristics of the regions, including a tight positive correlation between the amount of dust attenuation and the metal content. Although generalizing our results to whole galaxies may not be immediate, the possibility of a range of dust extinction laws and geometries should be accounted for in the latter systems as well. ", "introduction": "\\label{sec:introduction} Star formation is a fundamental process that governs the evolution of the baryonic matter in galaxies. It can deeply affect the host galaxy, its properties and appearance. The newly formed populations change the galaxy's spectral energy distribution (SED); the metals formed in the most massive stars pollute the interstellar medium and, in the presence of mechanical feedback from episodes of intense star formation, can be ejected from the galaxy and pollute the intergalactic medium. It is important to accurately measure star formation both as a function of spatial position within a galaxy and as a function of a galaxy's temporal evolution. The cosmic star formation rate density as a function of redshift \\citep{madau1996a,hopkins2006b} has become a classic comparison benchmark, and a challenge, for both semi-analytic and hydrodynamical models of structure formation and evolution \\citep[e.g.][]{katz1996a,hopkins2006a,stinson2006a,cattaneo2007a}. The youngest stellar populations, whose bolometric energy output is dominated by short-lived massive stars, emit the bulk of their energy in the restframe ultraviolet (UV). Since the pioneer work of \\cite{donas1984a} the UV has become one of the main star formation rate (SFR) tracers used in modern Astronomy \\citep{kennicutt1998a}. This regime, which has been previously difficult to observe in the local Universe, becomes the wavelength region ``par excellence'' to investigate star formation in galaxies at intermediate and high redshift. At these distances, the restframe UV is shifted to optical and near-infrared wavelengths, where current detectors afford the highest performance in terms of both sensitivity and angular resolution. Thus, it is no surprise that so far galaxy investigations in the redshift range z$\\sim$2-7 have been dominated by restframe UV observations \\citep{giavalisco2004a,bouwens2005a,mobasher2005a,sawicki2006a}. UV observations are the main canvass from which our current understanding of cosmic star formation is built. For over a decade now, studies of galaxies at cosmological distances have employed the ``starburst attenuation curve'' \\citep{calzetti1994a,calzetti2000a,meurer1999a} -- see next paragraph for a description of the physical conditions leading to such a law and equation 4 from \\cite{calzetti2000a} -- to correct their restframe UV measurements for the effects of dust extinction \\citep[e.g.][to name a few]{steidel1996a,steidel1999a,giavalisco2004a,daddi2005a,daddi2007a}. This curve is a powerful tool, in that it only requires information on the UV colors (or UV spectral slope) of a galaxy to provide an extinction correction, and recover the intrinsic UV flux. Assuming this curve, the typical UV extinction correction at redshift 3 is a factor $\\sim5$ in luminosity \\citep{steidel1999a}. However, in recent years, evidence has been accumulating that the starburst attenuation curve is not applicable to all galaxies \\citep{buat2002a,buat2005a,noll2009a} and that the UV extinction is dependent on the age of the stellar populations \\citep{cortese2008a}. More quiescently star-forming galaxies show characteristics in their UV SED that do not lend themselves to the same extinction ``treatment'' as more active starbursts. Applying the starburst attenuation curve to these more quiescent galaxies leads to up to an order of magnitude overestimate of the UV luminosity (and of the SFR). This result clearly represents a problem as deeper observations probe past the most powerful starburst to more normal galaxies with relatively lower SFR at increasingly higher redshifts. Along these lines, current determinations of UV luminosity functions at cosmological distances may be viewed with some suspicion, as all of them need some form of extinction correction \\citep{steidel1999a,giavalisco2004a,sawicki2006a,teplitz2006a}. One of the prescriptions of the starburst attenuation curve is based on a very simple, fully-empirical result. The ratio L(IR)/L(UV) is strongly correlated with the measured UV spectral slope $\\beta$ (or the UV color) in local, UV-bright starburst galaxies \\citep{meurer1999a,calzetti2000a}. The L(IR)/L(UV) ratio is a measure of the total dust opacity that affects the UV emission in a galaxy, or more generally, in a stellar system; the UV light absorbed by dust is re-emitted in the IR, independently of the details of the star-dust distribution, the dust properties, the metallicity, the stellar initial mass function, or the star formation history of the system \\citep{gordon2000a,buat2005a}. Meurer et al.'s result, commonly called the IRX-$\\beta$ relation, together with a host of previous results that showed strong correlations between $\\beta$ and a number of dust reddening measures, form the basis for an extinction correction prescription that obviously applies to local (and possibly distant) starburst galaxies \\citep{calzetti1994a,calzetti1996a,calzetti1997a,calzetti2001a}. The properties of the starburst attenuation curve imply specific conditions for the distribution of dust and stars within the starburst region \\citep{calzetti1994a,calzetti1996a,witt2000a,charlot2000a,calzetti2001a}. The geometry that can account for all observed properties at UV, optical, near-IR, and far-IR wavelengths is that of a dust distribution foreground to the stellar population responsible for the bulk of the UV emission. The term foreground is here loosely used to indicate any geometry that puts the stars in the center of a shell-like dust distribution, and does not make any inference on the dust distribution (which can be homogeneous or clumpy). Starbursts provide a mechanism for creating such specific dust-star geometry: massive star wind and supernova feedback can create expanding bubbles of hot gas that sweep away a significant fraction of the dust and gas that resides in the starburst site itself. Within this scenario, the gas and dust end up at the ``edges'' of the starburst site thus creating an inhomogeneously distributed ``shell''. In addition, the starbust extinction curve also requires a specific type of dust grains lacking the 2175\\AA\\ bump \\citep{gordon1997a} like that found in the SMC bar \\citep{gordon1998a,gordon2003a}. The starburst attenuation curve may thus not apply to a generic galaxy, as the coherent feedback mechanism present in starbursts may be missing in other galaxies. A number of studies have confirmed that more quiescent star-forming galaxies, although they generally trace a correlation in the IRX-$\\beta$ plane, are also up to an order of magnitude less obscured than starbursts for similar $\\beta$ values, and can have a larger dispersion, with a factor up to $\\sim$5 larger spread around the mean correlation \\citep{buat2002a,buat2005a,bell2002a,kong2004a,gordon2004a,seibert2005a,calzetti2005a,boissier2007a,dale2007a,munoz2009a}. UltraLuminous Infrared Galaxies (ULIRGs) also deviate from the starburst IRX-$\\beta$ correlation \\citep{goldader2002a}, and are located above it, i.e., opposite to the star-forming galaxies. ULIRGs are dustier, more compact, and denser systems than the UV-bright starbursts; therefore, their location in the IRX-$\\beta$ diagram is plausibly accounted for by an ``homogeneous'' mixture of dust and stars \\citep{calzetti2001a}. ULIRGs likely represent less of a concern for the use of the UV emission as a SFR tracer, as in general they will not be detectable at restframe UV wavelengths in cosmological samples. Quiescent star-forming galaxies are a far larger concern, as they are relatively bright in the UV and become increasingly detected as surveys go deeper and, by the nature of the luminosity function, are more numerous than starbursts. Alas, for these galaxies, the next step in the investigation, understanding and quantifying the nature of the deviation from the starbursts' IRX-$\\beta$ correlation, has so far eluded any attempt. \\cite{bell2002a} suggested that a combination of variations in the dust geometry and the characteristics (age, etc.) of the stellar populations could be the reason for the deviation observed in the HII regions of the Large Magellanic Cloud. The sparse data available, however, prevented further insights into the problem. Similar limitations were encountered by \\cite{gordon2004a} in their analysis of the star-forming regions in M~81. \\cite{kong2004a}, using a non-homogeneous UV dataset, suggested that the deviations of a large sample of quiescent star-forming galaxies from the starburst relation could be parametrized by the b-parameter, the ratio of the current to past average star formation rate. However, neither \\cite{seibert2005a}, using the more homogeneous GALEX All-sky Imaging Survey sample plus IR data from IRAS, nor \\cite{johnson2007b}, using SDSS data, Spitzer, and spectroscopic observations, confirmed Kong et al.'s suggestion. Some of these works have also produced ``mean'' obscuration curves for quiescent star-forming galaxies, but the intrinsic spread of the data around these mean curves is more than an order of magnitude, versus $\\sim$2-3 typical of the starbursts. This suggests the existence of a ``second parameter'' for the star-forming galaxies \\citep{kong2004a}, which needs to be identified and understood if the UV is going to be used as a tracer of SFR. In this paper, we investigate the possible reason for the deviation from the starburst IRX-$\\beta$ relation concentrating on sub-kpc star forming regions within galaxies, which provide simpler samples of stellar population than whole galaxies. We use GALEX observations to measure the UV luminosity, Spitzer ones to derive the total infrared luminosity and optical ones to estimate the age. In section 2 we present the sample of galaxies selected for this study, in section 3 we present the UV, IR and optical observations, in section 4 we present the data processing and photometry methods carried out, in section 5 we present how we model the star forming regions, in section 6 we proceed on to test the validity of the age as the ``second parameter'', in section 7 we test different extinction laws, in section 8 we test the metallicity, we discuss the results presented in section 9 and we finally conclude in section 10. ", "conclusions": "We have performed an analysis of {\\bf over 300} star forming regions in 8 local quiescent star forming galaxies to garner new insights on the IRX-$\\beta$ relation by using Spitzer infrared, GALEX ultraviolet and ground based optical broad band and narrow-band H$\\alpha$ data. Specifically we investigate the deviations of star forming regions in normal star forming galaxies from the locus identified by the starburst attenuation curve, in hope of obtaining information on similar deviations by normal star forming galaxies. The mean age of the stellar population, as traced by the $U-B$ color and by the equivalent width of H$\\alpha$, is shown to be relatively uncorrelated or only weakly correlated with the perpendicular distance of the data from the starburst attenuation curve in the IRX-beta diagram. Stellar population ages have been suggested as a potential culprit for the deviations in the case of normal star-forming galaxies \\citep[e.g.][via the b-parameter]{kong2004a}, but we do not find evidence for such trend in HII knots of galaxies. Our results are in agreement with similar results, obtained for whole galaxies, by \\cite{seibert2005a} and \\cite{johnson2007b}, but disagree with recent findings by Dale et al. (2009), also derived for whole galaxies. We find that a range of dust extinction/attenuation curves and dust/star geometries are actually required to account for the location on the IRX-beta plot of most of our HII knots. Specifically, we find that the starburst attenuation curve and the SMC extinction with foreground geometry bracket the location of most of the data points, not only on the IRX-beta diagram but on other plots reporting colors and other properties of the regions. For the reddest (in UV color) regions, we need to include contributions to the observed colors and fluxes of underlying older ($\\sim$ 300 Myr) stellar populations to account for the observed UV and $U-B$ colors. The mass ratio between the young (ionizing) stellar population and the older one is about 1/5. Individual star forming regions and integrated galaxies do not populate the same locus in the IRX-$\\beta$ diagram. Indeed, the former have a bluer UV color and are more tightly correlated. The most likely explanation is that older stellar populations up to $1-2\\times10^9$ years old, which have a redder UV color redden the observed UV color of star forming galaxies. To accurately determine the star formation rate, it is crucial to take into account this older population. Finally, a strong correlation is found with the metallicity, in the sense that the more metal rich regions tend to display redder UV colors and larger IR/UV values, as expected metallicity and dust content correlate with each other." }, "0910/0910.5221_arXiv.txt": { "abstract": "We calculate how the relic density of dark matter particles is altered when their annihilation is enhanced by the Sommerfeld mechanism due to a Yukawa interaction between the annihilating particles. Maintaining a dark matter abundance consistent with current observational bounds requires the normalization of the s-wave annihilation cross section to be decreased compared to a model without enhancement. The level of suppression depends on the specific parameters of the particle model, with the kinetic decoupling temperature having the most effect. We find that the cross section can be reduced by as much as an order of magnitude for extreme cases. We also compute the $\\mu-$type distortion of the CMB energy spectrum caused by energy injection from such Sommerfeld-enhanced annihilation. Our results indicate that in the vicinity of resonances, associated with bound states, distortions can be large enough to be excluded by the upper limit $\\vert\\mu\\vert\\leq9.0\\times10^{-5}$ found by the COBE/FIRAS experiment. ", "introduction": "If dark matter annihilates, the byproducts of the annihilation (positrons, neutrinos, gamma-rays, etc.) can leave non-gravitational signatures that, if observed, would be crucial for clarifying the nature of dark matter. In recent years, a number of observations have highlighted anomalies that might be explained by invoking dark matter annihilation in our Galactic halo. Among them are: (i) an anomalous abundance of positrons in cosmic rays above $10$~GeV according to the PAMELA experiment \\cite{Adriani-09} confirming and extending previous measurements by experiments such as AMS-01 \\cite{-07}; (ii) an excess of microwave emission from the galactic center as measured by the WMAP experiment, known as the ``WMAP haze'' \\cite{Hooper-Finkbeiner-Dobler-07} (iii) the apparent excess of diffuse galactic gamma-rays with energies above 1GeV inferred from observations by the EGRET satellite \\cite{deBoer-05}; (iv) analyses on the balloon experiments ATIC and PPB-BETS \\cite{Chang-08,Torii-08} have reported an excess in the total flux of electrons and positrons in cosmic rays. For the last two anomalies, we note that recent observations by FERMI seem inconsistent with the claimed gamma-ray excess in the EGRET data \\cite{Porter--09}, and that the excess in the electron and positron flux is smaller than previously thought \\cite{Abdo-09}. The latter is actually consistent with a modified cosmic ray propagation model that does not require additional primary sources of electrons and positrons \\cite{Grasso-09}. Although other astrophysical sources could explain these anomalies (see for example \\cite{Hooper-Blasi-DarioSerpico-09,Yuksel-Kistler-Stanev-09,Profumo-08} for an explanation for the PAMELA excess based on particle acceleration by pulsars, and \\cite{Fujita-09,Shaviv-Nakar-Piran-09} for one based on supernova remnants), dark matter annihilation offers an attractive solution. Large annihilation rates are needed, however, to explain the observations. In particular for the PAMELA data, the required annihilation rate is typically a few orders of magnitude larger than the value obtained by assuming the standard cross section inferred from the observed abundance of dark matter together with a smooth local distribution of dark matter \\cite{Cirelli-09}. Thus, an additional hypothesis is needed to boost the annihilation rate to the required levels and this must not change the present-day abundance of dark matter. Such a boost is difficult to obtain from the effects of substructures in the local dark matter distribution. Recent numerical simulations predict this to be remarkably smooth \\cite{Vogelsberger-09}. A detailed analysis of the impact of substructure on the production of positrons by \\cite{Lavalle-08} came to a similar conclusion. The possibility of a nearby ``spike'' of dark matter produced by an intermediate mass black hole seems also a priori implausible \\cite{Bringmann-Lavalle-Salati-09}. An alternative that has produced a plethora of papers in recent years is that of a Sommerfeld enhancement to the cross section produced by the mutual interaction of the annihilating dark matter particles. Such an interaction could be produced by a force carrier which might be any of the standard model weak force gauge bosons \\cite{Lattanzi-Silk-09} or a new force carrier \\cite{Arkani-Hamed-09}. For certain values of the parameters of these models, the enhancement is easily large enough to boost the cross section to the required values. However, a large cross section has a significant impact on other observables and may violate other constraints. For instance, \\cite{Kamionkowski-Profumo-08} showed that for the case where the cross section increases as $1/v$ (a particular case of Sommerfeld-enhancement models), where $v$ is the relative velocity of the annihilating particles, there are severe constraints from measurements of the diffuse extragalactic gamma-ray background radiation and from CMB constraints on ionization and heating of the intergalactic medium (IGM) by annihilation in the first generation of halos. In this scenario, the boost factors required to fit the above anomalies would be inconsistent with current constraints. This problem is avoided, however, by more general cases of the Sommerfeld enhancement where the effect saturates at low velocities \\cite{Arkani-Hamed-09,Lattanzi-Silk-09}. Recently, \\cite{Galli-09} and \\cite{Slatyer-Padmanabhan-Finkbeiner-09} analyzed constraints on the annihilation cross section from perturbations to the CMB angular power spectra resulting from heating and ionization of the photon-baryon plasma at recombination. They found interesting upper limits to models with Sommerfeld enhancement (see fig. 5 of \\cite{Galli-09}) that already rule out some extreme cases. At higher redshifts, the effects of the Sommerfeld enhancement have scarcely been treated. For example,the annihilation of dark matter impacts the predictions from big-bang nucleosynthesis on the abundance of light elements (e.g. \\cite{Jedamzik-04}). This could put constraints on certain models with Sommerfeld enhancement. However, this has only been studied in passing \\cite{Hisano-09,Jedamzik-Pospelov-09}. Also, the abundance of dark matter today is commonly assumed to be unaltered by this effect, apparently because the typical dark matter particle velocities at freeze-out are very large making the enhancement very close to one at that epoch \\cite{Arkani-Hamed-09,Kuhlen-Madau-Silk-09}. As we will show, this reasoning is flawed. The thermodynamic equilibrium between matter and radiation in the early Universe would be perturbed by energy released during a certain process, dark matter annihilation for example. This equilibrium tends to be restored by different interaction mechanisms: Compton scattering, double Compton emission and bremsstrahlung radiation. The efficiency of these to fully restore equilibrium varies with redshift. For $z\\gtrsim2\\times10^6$ they are efficient enough to restore distortions in the energy spectrum and thus, the photon distribution is that of a black body with a slightly higher temperature than the one in the case of no energy injection. For lower redshifts these mechanisms can not restore the black body spectrum. In particular, for $5.1\\times10^4\\lesssim z \\lesssim2\\times10^6$, the spectrum is perturbed into a Bose-Einstein distribution with a chemical potential $\\mu$ \\cite{Illarionov-Siuniaev-75}. In this paper, we revisit the impact of the Sommerfeld enhancement on the relic particle abundance and we show that its effect is not negligible. We also study, for the first time, $\\mu-$type distortions of the CMB spectrum due to energy deposition by dark matter annihilation in models with Sommerfeld enhancement. The paper is organized as follows. In section 2 we summarize the Sommerfeld enhancement and describe how we include it in our calculations. The relic density calculation is set out in detail in section 3. In section 4, the $\\mu-$type distortion to the CMB from annihilation with Sommerfeld enhancement is calculated. Finally we present a summary and our conclusions in section 5. ", "conclusions": "The prospects for dark matter detection have increased considerably in recent years. There is a continual advance in the sensitivity of detectors on Earth that look for direct dark matter elastic scattering with nuclei, and in experiments that search indirectly for dark mater by looking for non-gravitational signatures of the byproducts of its hypothetical annihilation. Such technological improvements may lead, in the near future, to a definite proof of the existence of dark matter. Perhaps dark matter has produced non-gravitational signals that have already been detected. The recently reported excesses of positrons in cosmic rays by PAMELA and of electrons+positrons by FERMI/ATIC/PPB-BETS can be explained by dark matter annihilation. Although other explanations with a different astrophysical origin are also possible, a solution based on dark matter is an attractive possibility. However, such solution seems to require an attractive force between the dark matter particles to enhance their annihilation through the Sommerfeld mechanism. For a Yukawa interaction via a single scalar, the magnitude of the enhancement depends on the coupling strength of the interaction $\\alpha_c$, on the mass ratio of the force carrier to the dark matter particle $m_{\\phi}/m_{\\chi}$ and on the relative velocities of the annihilation particles $\\beta$. Several recent papers have computed the boost to the annihilation due to a Sommerfeld enhancement (in the Galactic halo and/or in its subhalos) \\cite{Lattanzi-Silk-09,Arkani-Hamed-09,Bovy-09}. One of the main aims of these works is to show that for certain values of $\\alpha_c$ and $m_{\\phi}/m_{\\chi}$, this mechanism is able to produce large enough boosts to explain the cosmic ray anomalies. Only a handful of studies have addressed in detail the impact that such an enhancement has in the early Universe. In \\cite{Kamionkowski-Profumo-08}, the authors found that if the cross section increases with decreasing relative velocity as $1/\\beta$ (which is valid in a certain regime for general Sommerfeld enhancement models), dark matter annihilation in the first halos would heat and ionize the IGM, violating current constraints from the CMB. As was later pointed out by \\cite{Lattanzi-Silk-09} and \\cite{Arkani-Hamed-09}, this problem is alleviated in more general models where the enhancement saturates at low velocities. In \\cite{Galli-09} and \\cite{Slatyer-Padmanabhan-Finkbeiner-09}, a constraint on the annihilation cross section was obtained by considering limits on the energy deposition by annihilation at recombination. The constraint reported by \\cite{Slatyer-Padmanabhan-Finkbeiner-09} is $\\langle\\sigma v\\rangle_{REC}<3.6\\times10^{-24}{\\rm cm}^3{\\rm s}^{-1}(m_{\\chi}/1{\\rm TeV})/f_{REC}$, where $f_{REC}$ is an average efficiency of energy injection into the IGM by annihilation at recombination. In the present paper, we have analyzed the impact of dark matter annihilation with Sommerfeld enhancement at higher redshift. In the first place, we have revisited the calculation of the dark matter particle abundance by solving the Boltzmann equation from freeze-out through the epoch of kinetic decoupling, including a full solution to the Schr\\\"odinger equation. Contrary to previous claims \\cite{Arkani-Hamed-09,Kuhlen-Madau-Silk-09}, we have found a significant suppression of the relic density, in agreement with a recent work by \\cite{Dent-Dutta-Scherrer-09}. This suppression is particularly important near resonances, which are typically invoked to explain the cosmic ray anomalies. We found that to fit the observed dark matter abundance, the normalization of the cross section needs to be lowered by up to a factor of 10 compared to the case without enhancement (see Fig.~\\ref{cross_section}). The result depends on the coupling strength and the proximity to a resonance. Exploring a broad range of dark matter particle masses and kinetic decoupling temperatures, we found a minor-to-medium impact of these parameters on our results; variations on $T_{kd}$ have the strongest impact. Secondly, we have calculated the amount of energy deposited by dark matter annihilation into the radiation plasma in the redshift range $5.1\\times10^46\\times10^{-2}$ already ruled out (see Fig.~\\ref{mu_coulomb}). Improved upper limits on the $\\mu$ parameter by a null distortion detection in the CMB spectrum would rule out a larger region of the parameter space for the Yukawa interaction. In \\cite{Fixsen-Mather-02}, it was pointed out that an improvement close to two orders of magnitude has already been possible for a number of years, and \\cite{Mather-07} suggests that another order of magnitude is perhaps within reach. An upper limit on $\\vert\\mu\\vert$ of the order of $10^{-7}$ would certainly exclude large regions of the parameter space. It would exclude to a large extent near-resonance regions, which are the ones that produce the largest boosts to the annihilation. On the other hand, the detection of a distortion could possibly tell us something about the parameters of the Yukawa interaction and the nature of the force carrier $\\phi$. In summary, our results indicate that for a given set of parameters $(m_{\\chi},T_{kd},m_{\\phi},\\alpha_c)$, it is necessary to compute in detail the relic density to obtain the range of values for the normalization to the cross section that are compatible with current estimates of the dark matter abundance. Once this normalization is known, the energy input producing a $\\mu-$type distortion in the CMB spectrum can be computed to check whether the particular model violates current constraints. Only for allowed models can a boost factor for local dark matter annihilation be computed and advocated. Our findings show that the local boosts reported in the literature need to be renormalized to the proper value of $\\langle\\sigma v\\rangle_S$ implied by the observed relic density. This renormalization can exceed a factor of 10 in extreme cases. \\begin{figure} \\centering \\includegraphics[height=8.5cm,width=10cm]{./fig6.ps} \\caption{The values of the boost factor, relative to $\\langle\\sigma v\\rangle_S=3\\times10^{-26}{\\rm cm}^3{\\rm s}^{-1}$ for the parameter scan of the Yukawa interaction. These values are obtained according to the cross section values of Fig.~\\ref{cross_section}, which are consistent with $\\Omega_{\\chi}h^2=0.1143$. A Maxwell-Boltzmann velocity distribution with $\\sigma_v=5\\times10^{-4}$ for a neutralino with $m_{\\chi}=100$GeV and $T_{kd}=8\\times10^{-3}$GeV was used. The values are color-coded logarithmically according to the scale on the right.} \\label{boost_factor} \\end{figure} A recent analysis of the PAMELA and FERMI data by \\cite{Bergstrom-Edsjo-Zaharijas-09} suggests that boost factors, over a standard value of $\\langle\\sigma v\\rangle_S=3\\times10^{-26}{\\rm cm}^3{\\rm s}^{-1}$, of order 1000 or higher are required to explain both datasets simultaneously. We note that smaller boost factors ($\\sim 100$) can still fit the PAMELA data, but not the FERMI one, for a 100~GeV neutralino, which is the standard case we have considered here. These large boosts are also found in the analysis by \\cite{Meade-09}. It has been argued in the past that these boosts can be achieved only by invoking Sommerfeld-enhanced annihilation. For instance, \\cite{Arkani-Hamed-09} and \\cite{Bovy-09} obtain maximum boost factors $\\sim1000$ by assuming a local Maxwell-Boltzmann velocity distribution with a velocity dispersion of $150~{\\rm kms}^{-1}$ ($5\\times10^{-4}c$), which roughly corresponds to the estimated local dark matter velocity dispersion. We find that these boosts are modified once the proper cross section is used. In Fig.~\\ref{boost_factor} we show the boost factors, i.e., the multiplicative factor to $\\langle\\sigma v\\rangle_S=3\\times10^{-26}{\\rm cm}^3{\\rm s}^{-1}$, that we find for a Maxwell-Boltzmann velocity distribution with $\\sigma_v=5\\times10^{-4}$ for the parameter space of the Yukawa interaction. The figure is for $m_{\\chi}=100$GeV, $T_{kd}=8\\times10^{-3}$GeV and is consistent with a relic abundance $\\Omega_{\\chi}h^2=0.1143$. Keep in mind, however, that changes in the neutralino mass have almost no impact on the normalization of the cross section. Different kinetic decoupling temperatures and variations on $\\Omega_{\\chi}h^2$ have also only a small effect ($<60\\%$ in combination). Thus, Fig.~\\ref{boost_factor} is approximately correct for all the cases considered in this paper. The figure shows that even in extreme cases, for resonances in the upper right of the figure, the boost factors are $<500$. In more favored regions of the parameter space $\\alpha_c\\gtrsim10^{-2}$, $m_{\\phi}/m_{\\chi}\\lesssim10^{-3}$ (see for example \\cite{Bovy-09}), the boost factors are $\\lesssim100$. This result suggests that additional assumptions are needed to account for the boost needed to explain the cosmic ray anomalies by dark matter annihilation alone. The inclusion of colder substructures to the overall smooth component with higher densities and lower velocity dispersions and thus, higher boost factors, could perhaps solve the issue. However, as found in recent high-resolution N-body simulations, the local dark matter distribution is rather smooth \\cite{Vogelsberger-09}. Typical estimates on such an additional boost factor due to substructure in the galactic halo (albeit without Sommerfeld enhancement) are of the order of 1.4, unless a subhalo happens to be very close to Earth, in which case this boost could be larger than 10 \\cite{Diemand-08}. These estimates do not include a possible further amplification due to the scaling of the Sommerfeld enhancement with velocity dispersion. However, the constraints found by \\cite{Slatyer-Padmanabhan-Finkbeiner-09} based on energy deposition at recombination suggest that the enhancement must be already close to saturation for $\\sigma_v=150{\\rm kms}^{-1}$, thus such a further amplification seems implausible. Therefore, is not clear that the inclusion of colder substructures can account for the additional boost. Finally, we mention that we have created a web application that solves the relevant equations described in this paper. It allows the user to compute individual Sommerfeld boosts, thermally averaged enhancements, cross section and $\\mu$ values, among other quantities. The interested reader can find this at http://www.mpa-garching.mpg.de/$\\sim$vogelsma/sommerfeld/" }, "0910/0910.2681.txt": { "abstract": "We explore how the local environment is related to properties of active galactic nuclei (AGNs) of various luminosities. Recent simulations and observations are converging on the view that the extreme luminosity of quasars, the brightest of AGNs, is fueled in major mergers of gas-rich galaxies. In such a picture, quasars, the highest luminosity AGNs, are expected to be located in regions with a higher density of galaxies on small scales where mergers are more likely to take place. However, in this picture, the activity observed in low-luminosity AGNs is due to secular processes that are less dependent on the local galaxy density. To test this hypothesis, we compare the local photometric galaxy density on kiloparsec scales around spectroscopic type I and type II quasars to the local density around lower-luminosity spectroscopic type I and type II AGNs. To minimize projection effects and evolution in the photometric galaxy sample we use to characterize AGN environments, we place our random control sample at the same redshift as our AGNs and impose a narrow redshift window around both the AGNs and control targets. Our results support these merger models for bright AGN origins. We find that the brightest sources have overdensities that increase on the smallest scales compared to dimmer sources. In addition, we investigate the nature of the quasar and AGN environments themselves and find that the increased overdensity of early-type galaxies in the environments of bright type I sources suggests that they are located in richer cluster environments than dim sources. We measure increased environment overdensity with increased quasar black hole mass, consistent with the well-known $M_{\\rm DMH}-\\MBH$ relationship, and find evidence for quenching in the environments of high accretion efficiency type I quasars. ", "introduction": " ", "conclusions": "" }, "0910/0910.0632_arXiv.txt": { "abstract": "{% The Square Kilometre Array (SKA) is intended as the next-generation radio telescope and will address fundamental questions in astrophysics, physics, and astrobiology. The international science community has developed a set of Key Science Programs: (1)~Emerging from the Dark Ages and the Epoch of Reionization, (2)~Galaxy Evolution, Cosmology, and Dark Energy, (3)~The Origin and Evolution of Cosmic Magnetism, (4)~Strong Field Tests of Gravity Using Pulsars and Black Holes, and (5)~The Cradle of Life/Astrobiology. In addition, there is a design philosophy of ``exploration of the unknown,'' in which the objective is to keep the design as flexible as possible to allow for future discoveries. Both a significant challenge and opportunity for the \\hbox{SKA} is to obtain a significantly wider field of view than has been obtained with radio telescopes traditionally. Given the breadth of coverage of cosmic magnetism and galaxy evolution in this conference, I highlight some of the opportunities that an expanded field of view will present for other Key Science Programs.} \\FullConference{Panoramic Radio Astronomy: Wide-field 1-2 GHz research on galaxy evolution - PRA2009\\\\ June 02 - 05 2009\\\\ Groningen, the Netherlands} \\begin{document} ", "introduction": "\\label{sec:jl.intro} In the $20^{\\mathrm{th}}$ Century, we discovered our place in the Universe. We learned that it was much bigger than we imagined and much more exotic. Beyond our Milky Way, the Universe is filled with galaxies---each their own island universe. They range in size from dwarf galaxies barely able to survive near their larger neighbors to giant elliptical galaxies, orders of magnitudes larger than the Milky Way. These galaxies of stars also contain a multitude of other components including gas with a wide range of temperatures; compact objects including white dwarfs, neutron stars, and black holes; and planets. Over the course of the century, black holes moved from a theoretical curiosity to a well-recognized endpoint of stellar evolution and a likely fundamental component of the centers of galaxies, with the potential to power immense jets of relativistic particles that affect their surroundings. By the end of the century, we were beginning to unveil the basic structure and processes of the Universe in which these objects are embedded, including evidence of its origin and the still mysterious properties of dark matter and dark energy. Our probes of the Universe have expanded dramatically as well. Electromagnetic radiation has been detected from celestial objects at frequencies below~1~MHz ($\\lambda \\sim 300$~m) to energies exceeding 1~\\hbox{TeV}. Moreover, the range of possible signals has expanded beyond just electromagnetic radiation. Cosmic rays rain down on the Earth, some with energies approaching those of macroscopic objects. Gravitational radiation has been detected indirectly, and numerous potential classes of sources have been suggested, with the expectation that the Earth is awash in gravitational waves. Neutrinos have been detected from both the Sun and supernova \\hbox{1987A}, and many of the processes that generate high energy cosmic rays should also produce a spectrum of high-energy neutrinos. In the $21^{\\mathrm{st}}$ Century, we seek to understand the Universe we inhabit. To do so will require a suite of powerful new instruments, on the ground and in space, operating across the entire electromagnetic spectrum and for multiple decades. Observations at centimeter- to meter wavelengths have provided deep insight to a wide range of phenomena ranging from the solar system to the most distant observable celestial emission. This long and rich record of important discoveries in the radio spectrum, including 3 Nobel prizes, has been possible since many of the relevant physical phenomena can only be observed, or understood best, at these wavelengths. These phenomena include the cosmic microwave background (CMB), quasars, pulsars, gravitational waves, astrophysical masers, magnetism from planets through galaxies, the ubiquitous jets from black holes and other objects, and the spatial distribution of hydrogen gas, the predominant baryonic constituent of the Universe. Moreover, through the invention of aperture synthesis, also recognized by the Nobel committee, radio astronomy has reached unprecedented levels of imaging resolution and astrometric precision, providing the fuel for further discovery. With only a handful of exceptions, radio telescopes and arrays have been limited to apertures of about~$10^4$~m${}^2$, constraining, for instance, studies of the 21-centimeter hydrogen emission to the nearby Universe ($z \\sim 0.2$)~\\cite{cr04}. Contemporaneous with the astronomical discoveries in the latter half of the $20^{\\mathrm{th}}$ Century have been technological developments that offer a path to substantial improvements in future radio astronomical measurements. Among the improvements are mass production of centimeter-wavelength antennas enabling apertures potentially 100 times larger than previously available, fiber optics for the transmission of large volumes of data, high-speed digital signal processing hardware for the acquisition and analysis of the signals, and computational improvements leading to massive processing and storage. These new technologies, combined with dramatically improved survey speeds and the other advances, can open up an enormous expanded volume of discovery space, providing access to many new celestial phenomena and structures, including 3-dimensional mapping of the web of hydrogen gas through much of cosmic history ($z \\sim 2$). The realization that radio astronomy was on the doorstep of a revolutionary age of scientific breakthrough has led the international community to investigate this opportunity in great detail over the last decade. That coordinated effort, involving a significant fraction of the world's radio astronomers and engineers, has resulted in the Square Kilometre Array (SKA) Program (Figure~\\ref{fig:ska}), an international roadmap for the future of radio astronomy over the next two decades and one for which access to a wide field of view is an integral part of the science. \\begin{figure} \\begin{center} \\includegraphics[width=0.9\\textwidth]{SKA_cores_2009.eps} \\end{center} \\vspace*{-3ex} \\caption{An artist's impression of the core of the SKA illustrating the various technologies over the frequency range 70~MHz to~10~GHz. All of these technologies would enable various levels of wide-field imaging.} \\label{fig:ska} \\end{figure} From its inception, development of the SKA Program has been a global endeavor. In the early 1990s, there were multiple, independent suggestions for a ``large hydrogen telescope.'' It was recognized that probing the fundamental baryonic component of the Universe much beyond the local Universe would require a substantial increase in collecting area. The IAU established a working group in~1993 to begin a worldwide study of the next generation radio observatory. Since that time, the effort has grown to comprise 19 countries and more than 50 institutes, including about~200 scientists and engineers. ", "conclusions": "" }, "0910/0910.0404_arXiv.txt": { "abstract": "The kinematics of the extra-planar neutral and ionised gas in disc galaxies shows a systematic decline of the rotational velocity with height from the plane (vertical gradient). This feature is not expected for a barotropic gas, whilst it is well reproduced by baroclinic fluid homogeneous models. The problem with the latter is that they require gas temperatures (above $10^5$ K) much higher than the temperatures of the cold and warm components of the extra-planar gas layer. In this paper, we attempt to overcome this problem by describing the extra-planar gas as a system of gas clouds obeying the Jeans equations. In particular, we consider models having the observed extra-planar gas distribution and gravitational potential of the disc galaxy NGC\\,891: for each model we construct pseudo-data cubes and we compare them with the HI data cube of NGC\\,891. In all cases the rotational velocity gradients are in qualitative agreement with the observations, but the synthetic and the observed data cubes of NGC\\,891 show systematic differences that cannot be accommodated by any of the explored models. We conclude that the extra-planar gas in disc galaxies cannot be satisfactorily described by a stationary Jeans-like system of gas clouds. ", "introduction": "Observations of several spiral galaxies at various wavelengths have revealed massive gaseous haloes surrounding the galactic discs. This extra-planar gas is multiphase: it is detected in \\ion{H}{I} (e.g. Swaters, Sancisi \\& van der Hulst 1997), in H$\\alpha$ (Rand 2000; Rossa et al. 2004) and in X-ray observations (Wang et al. 2001; Strickland et al. 2004). Thanks to high-sensitivity \\ion{H}{I} observations of some edge-on galaxies, like NGC 891 (Oosterloo, Fraternali \\& Sancisi 2007), we can trace these haloes up to $10$ - $20$ kpc from the plane and we can study their kinematics in detail. A remarkable feature of the extra-planar gas is its regularly decreasing rotational velocity $u_{\\varphi}$ at increasing distances from the galactic plane (vertical gradient). Fraternali et al.\\ (2005) found that in NGC 891 the vertical gradient is $\\sim\\, -15 {\\rm km \\,s} ^{-1} {\\rm kpc}^{-1}$ (see also Heald, Rand \\& Benjamin et al. 2007). For other galaxies (e.g. NGC 2403) gradients of the same order have been estimated (see Fraternali 2008). The dynamical state of the extra-planar gas can provide useful insights to understand its origin. Two main frameworks have been proposed: the galactic fountain (Shapiro \\& Field 1976; Bregman 1980) and the cosmological accretion (Oort 1970; Binney 2005). In the galactic fountain model, partially ionized gas is ejected from the disc by supernova explosions and stellar winds, it travels through the halo and eventually falls back to the disc. Due to difficulties in running high-resolution hydrodynamical simulations of the whole galactic disc (see e.g.\\ Melioli et al.\\ 2008), the gas in the galactic fountain model is often treated ballistically, i.e. as a collection of non-interacting clouds subject to the gravitational field of the galaxy (Collins, Benjamin \\& Rand 2002). The ballistic galactic fountain is able to reproduce the gas distribution observed in spiral galaxies, but in general the predicted vertical gradient of $u_{\\varphi}$ is significantly lower than observed (Fraternali \\& Binney 2006). In the accretion models, the extra-planar gas is the result \\rev{of the infall of cool intergalactic gas}. The two scenarios are not mutually exclusive: in fact it has been argued that the contribution of accretion is limited to only $10-20\\%$ of the total mass of extra-planar gas (Fraternali \\& Binney 2008). In a complementary approach, some authors (Benjamin 2002; Barnab\\`{e} et al. 2006), have focused on the construction of models for the extra-planar gas under the assumption of stationarity. In this approach the problem of the origin of the extra-planar gas is not addressed, and the main effort is to clarify the global dynamical state of the gas. Usually, in the stationary approach, the extra-planar gas is described as an homogeneous fluid in permanent rotation. In the simplest of fluid homogeneous models, vertical and radial motions of the gas are neglected and therefore the vertical gravitational field of the galaxy is balanced only by the pressure gradient of the gas. In other studies, turbulence has also been added as an extra-pressure term (e.g., Koyama \\& Ostriker 2008). According to the Poincar\\'{e}-Wavre theorem (Lebovitz 1967; Tassoul 1978), a vertical gradient in the rotational velocity can be present only if the pressure of the medium is not stratified on the density, therefore in fluid homogeneous models the gas distribution is necessarily \\textit{baroclinic} (Barnab\\`{e} et al. 2006; see also Waxman 1978). Remarkably, in their study of NGC 891, Barnab\\`{e} et al. (2006) found that if the baroclinic configuration is built from general physical arguments, the model rotational velocity well reproduces the observed rotation curves at different heights over the galactic disc and for a large radial range. However, the temperature predicted for the system is $> 10^{5}$ K, well above that of the neutral gas. Here we investigate if and how this ``temperature problem'' can be solved. We make use of the fact that stationary fluid equations for a gaseous system in permanent rotation are, from a formal point of view, identical to the stationary Jeans equations for an axisymmetric system with isotropic velocity dispersion. Thus, also in this case, a decrease of rotational velocities with increasing distances from the mid-plane is obtained when the density field is built using the approach described by Barnab\\`{e} et al. (2006). In this paper we analyse this alternative interpretation of the baroclinic fluid homogeneous models. In practice we consider a ``gas'' of cold \\ion{H}{I} clouds described by the stationary Jeans equations, where the thermal pressure, needed for the vertical balance of the galaxy gravitational field, is replaced by the velocity dispersion of the clouds. \\rev{ In the present approach, some preliminary consideration on the physical state of the clouds is important. In fact, \\ion{H}{I} halo clouds are often assumed to be embedded in, and in pressure equilibrium with, a hot medium (corona) that is about a thousand times less dense and provides the pressure required to confine them (Spitzer 1956; Wolfire et al. 1995). In the Milky Way, there is plenty of evidence for the existence of this medium, such as emission from highly ionised metals (e.g. Sembach et al.\\ 2003) and the head-tail morphology of individual High Velocity Clouds (HVCs, e.g. Br{\\\"u}ns et al. 2000). Due to the high density contrast, the pressure-confined clouds cannot be supported by the pressure of the external medium and must be moving more or less ballistically in a way akin to water drops in the air (Bregman 1980). This assumption is commonly employed in galactic fountain models (e.g. Collins et al. 2002). On the time-scale taken for a cloud to move about 1000 times its length, the trajectory may deviate significantly from the ballistic one due to the interactions with the external medium. This time-scale depends on the cloud mass but it is longer than a dynamical time for the larger clouds (Fraternali \\& Binney 2008). The typical mass of an halo cloud is difficult to estimate. In the Milky Way the new determinations of distances for the HVCs give masses from a few $\\times 10^4 \\mo$ to $10^6 \\mo$ (e.g. Wakker et al. 2008), for external galaxies the mass resolution is often above $10^6 \\mo$ but in M31 several clouds with masses down to $\\sim 1 \\times 10^5 \\mo$ have been observed (Thilker et al. 2004). We estimate that for a cloud of $10^5 \\mo$, the drag time scale is more than 10 times the dynamical time and it is therefore fair to treat the system as purely dynamical. Moreover, if the \\ion{H}{I} halo of a galaxy like NGC\\,891 is made up of clouds with this mass and radii of $100$ pc, the collision time between clouds turns out to be about 5 times the dynamical time and the system can be considered collision-less. } The substitution of a fluid system with a gas of clouds is also not trivial from the point of view of the comparison with the observations, as several delicate issues arise when considering \\ion{H}{I} observations (see Sect. \\ref{constr}). To overcome these problems we follow Fraternali \\& Binney (2006, 2008) and construct, as an output for each of the models investigated, a \\textit{pseudo-data cube} with the same resolution and total flux as the \\textit{data cube} of the \\ion{H}{I} observations. This procedure assures a full control of projection and resolution effects. Moreover, the comparison with the raw data removes all the intermediate stages of data analysis and the associated uncertainties. For completeness, we also investigate some phenomenological anisotropic models. The paper is organized as follows: in Section 2 we illustrate the Jeans-based interpretation of the baroclinic solutions, introducing the anisotropic case which has no analogous fluid counterpart, and we discuss the conditions required for these solutions to have a negative vertical velocity gradient. In Section~3 we present the method adopted to construct isotropic and anisotropic models for the edge-on galaxy NGC 891, whilst in Section~4 we compare their predictions with the \\ion{H}{I} observations. Section~5 is devoted to the discussion of the results, and Section~6 concludes. ", "conclusions": "In this paper, motivated by the results of Barnab\\`e et al. (2006), we investigated the possibility that the extra-planar gas in spiral galaxies can be modelled as a ``gas'' composed by cold \\ion{H}{I} clouds that follows the stationary Jeans equations. \\rev{In doing this we are assuming that the pressure-bound clouds are moving almost ballistically in the halo. This assumption is valid in the limit of massive clouds having negligible rates of mass exchange with the corona. } In this alternative interpretation of fluid baroclinic models the thermal pressure is replaced by an isotropic velocity dispersion tensor, so that the problem of high temperature of gaseous homogeneous models is eliminated. We have also extended the discussion to simple phenomenological models with anisotropic velocity dispersion tensors. We constructed both isotropic and anisotropic models in the well-constrained gravitational field of the spiral galaxy NGC 891, and compared their predictions to the observed kinematics of the extra-planar gas in that galaxy. For each model we built a pseudo-data cube with the same resolution and total flux as the observations. The main results of our analysis can be summarized as follows: \\begin{enumerate} \\item The adopted functional form of the cloud density distribution, taken by the observations, leads to physically acceptable solutions in all the models investigated. The cloud density distribution is centrally depressed, and, in order to match the vertical extension of the \\ion{H}{I} halo of NGC 891, it has higher scale-height than that used by Barnab\\`{e} et al. (2006). \\item All the models computed show a negative vertical gradient in the rotational velocity, the distinctive feature of the kinematics of the extra-planar gas. \\item The support against the vertical gravitational field of the galaxy requires a $\\sigma_{z} \\simeq 50 - 100\\, \\rm{km\\, s^{-1}}$ and therefore the line-of-sight velocity dispersion of the isotropic model is a factor $\\sim\\,3-4$ higher than that observed. \\item With the introduction of the anisotropy it is possible to restrict the line-of-sight velocity dispersion to the observed values, but the predicted vertical gradient in the rotational velocity is somewhat too shallow and other features of the data cube are not fully reproduced. \\end{enumerate} We conclude that the dynamics of extra-planar gas in a galaxy like NGC\\,891 is not fully described by any of the stationary models considered here. However a model with an anisotropic velocity dispersion tensor, which mimics a galactic fountain is the preferable among all. \\rev{The fact that none of the stationary Jeans models analysed here can reproduce all the features of the observed (extra-planar) gas kinematics might suggest that the cloud motion is not purely ballistic, and that the interaction between the clouds and the coronal gas, perhaps in the form of mass exchange, plays an important dynamical role.}" }, "0910/0910.4011_arXiv.txt": { "abstract": "We present $25''$ resolution radio images of five Lynds Dark Nebulae (L675, L944, L1103, L1111 and L1246) at 16\\,GHz made with the Arcminute Microkelvin Imager (AMI) Large Array. These objects were previously observed with the AMI Small Array to have an excess of emission at microwave frequencies relative to lower frequency radio data. In L675 we find a flat spectrum compact radio counterpart to the 850\\,$\\mu$m emission seen with SCUBA and suggest that it is cm-wave emission from a previously unknown deeply embedded young protostar. In the case of L1246 the cm-wave emission is spatially correlated with 8\\,$\\mu$m emission seen with \\emph{Spitzer}. Since the MIR emission is present only in \\emph{Spitzer} band 4 we suggest that it arises from a population of PAH molecules, which also give rise to the cm-wave emission through spinning dust emission. ", "introduction": "The complete characterization of microwave emission from spinning dust grains is a key question in both astrophysics and cosmology. It probes a region of the electromagnetic spectrum where a number of different astrophysical disciplines overlap. It is important for CMB observations in order to correctly characterise the contaminating foreground emission; for star and planetary formation it is important because it potentially probes a regime of grain sizes that is not otherwise easily observable. Although a number of objects have now been found to exhibit anomalous microwave emission, attributed to spinning dust, it is still unclear what differentiates those objects from the many other seemingly similar targets that do not show the excess. In the specific case of dark clouds the recent AMI sample (AMI Consortium: Scaife et~al.\\ 2009; hereinafter Paper I) of fourteen Lynds Dark Nebulae found an excess in only five. It has been suggested that cm-wave emission from spinning dust is emitted by a population of ultra-small grains (Draine \\& Lazarian 1998). These ultra-small grains are thought to exist mainly in the form of single polycyclic aromatic hydrocarbon (PAH) molecules. PAH molecules are generally detected through their narrow line emission features in the MIR. For these emission features to be observed the PAH molecules must be exposed to a strong source of UV flux. Since this flux is generally absent in the case of dark clouds, the microwave emission from the rotation of PAH molecules may be the only way to study the very small grain population in these objects. It is also known that radio continuum emission in dark clouds may arise from ionized gas associated with a stellar outflow. When a luminous star is present this arises either as the result of a compact {\\sc{Hii}} region or an ionized stellar wind. In the case of very young low luminosity stars radio continuum emission may be also be detected. In this instance it is generally attributed to the presence of a partially ionized ($0.02\\leq x_{\\rm e} \\leq 0.35$; Bacciotti \\& Eisl{\\\"o}ffel 1999) stellar wind (Wright \\& Barlow 1975; Panagia \\& Felli 1975), or possibly a neutral wind which has been shock-ionized further from the central source by impacting on a dense obstacle (Curiel et~al.\\ 1989). In this paper we present follow-up observations of the five AMI Small Array (SA) spinning dust detections (Paper I) at higher resolution with the AMI Large Array (LA) over the same frequency range. All co-ordinates in this paper are J2000.0. ", "conclusions": "L675 and L1246 have archival \\emph{Spitzer} IRAC data, which shows in both cases a significant amount of emission in Band 4 (6.4--9.4\\,$\\mu$m) and very little in the other three (3.2--3.9, 4.0--5.0 and 4.9--6.4\\,$\\mu$m, respectively). In the case of L675 this emission is present on a very large scale, see Fig.~\\ref{fig:l675spit}. The emission seen at 16\\,GHz with the AMI SA appears on a similar scale, however the small field of view of the \\emph{Spitzer} data precludes a more detailed comparison. L1246 shows an arc of emission at 16\\,GHz which is also evident in \\emph{Spitzer} IRAC Band 4, see Fig.~\\ref{fig:l1246spit}. This emission is again not present in Bands 1--3. In Band 4 it is present as an arc, coincident with that seen at 16\\,GHz in the AMI LA data. \\emph{Spitzer} Band 4 contains two of the PAH emission lines, including the strongest (7.7\\,$\\mu$m). Of the three other \\emph{Spitzer} bands only Band 1 contains an emission line (3.3\\,$\\mu$m) and for ionized PAHs this line is expected to be significantly weaker. It is probable therefore that the MIR correlated cm-wave data seen in the AMI maps is a consequence of spinning dust emission from a population of ionized PAH molecules. Neutral PAH molecules do not in general possess a permanent dipole moment and are therefore not expected to have rotational emission (Tielens 2008). This emission, the mechanism of which is described in detail by Draine \\& Lazarian (1998), arises from the intrinsic dipole moments of small dust grains, most likely to be PAH molecules, which emit power when they rotate. This rotation has a variety of contributing factors, the relative importance of which varies with grain environment. However, in the majority of cases excitation through collision with ions is predominant. In the case of L675A, we must consider the possibility that we are observing a coincidental extragalactic radio source. Using the extended 9C survey 15\\,GHz source counts (Waldram et~al.\\ 2009), where $n(S) = 51 (S/{\\rm{Jy}})^{-2.15}$\\,Jy$^{-1}$\\,sr$^{-1}$, the probability that a source with flux density greater than 2\\,mJy lies within the FWHM of the AMI LA primary beam is $0.12$, and only 0.01 within the SCUBA field. It is likely therefore that the radio source L675A is associated with the SCUBA core. A further question is whether the cm-wave emission might be explained by thermal (Planckian) dust emission. A single greybody spectrum with a dust temperature, $T_{\\rm{d}} \\approx 27$\\,K, might be used to explain the LA flux density, however it would require a $\\beta$ of 0.6. Such a value would be unusual even for objects known to possess flattened dust tails, such as protoplanetary disks. This simple fit also neglects the flux lost by the AMI LA baseline distribution. SA observations have already shown this source to possess a significant amount of extended emission which would make this scenario even more unlikely. The presence of a neutral or partially ionized wind from an outflow source that has been shocked through encountering a dense obstacle (Torrelles et~al.\\ 1985; Rodr{\\'i}guez et~al.\\ 1986) is used to understand the spectral indices seen towards exciting sources in the radio regime (Curiel et~al.\\ 1990; Cabrit \\& Bertout 1992). This model allows a spectral index range of $0.1$ (optically thin) to -2 (optically thick), which explains results which deviate from the value of $\\alpha = -0.6$ required by a spherically symmetric ionized wind (Wright \\& Barlow 1975; Panagia \\& Felli 1975). Using this model as described in Curiel et~al.\\ (1989; 1990) the radio emission is expected to be optically thin ($\\tau=0.1$), consistent with the spectral index seen across the AMI band. Assuming a distance of 300\\,pc and a stellar wind with a wind speed of 200\\,km\\,s$^{-1}$, we can calculate that the AMI flux densities towards L675A are consistent with a mass loss of $3.5\\times 10^{-7}$\\,M$_{\\odot}$\\,yr$^{-1}$. A mass loss such as this implies a mechanical luminosity from the wind of $L_{\\rm{mech}} \\approx 1.1$\\,L$_{\\odot}$, comparable to the values found by Curiel et~al for L1448. The nature of the emission seen towards L944 with the AMI LA is uncertain. The spectral index of this emission is consistent with spinning dust emission or alternatively the optically thick component of free--free spectrum. Such a free--free spectrum might be exhibited at 16\\,GHz by ultra-compact {\\sc Hii} regions. However a turn-over frequency above 16\\,GHz would have an extremely high mass and should therefore be obvious in sub-mm observations. This needs to be confirmed by either higher radio frequency measurements in order to measure the optically thin region of the spectrum and the turn-over, or sub-mm measurements to place constraints on the mass of such a region. In conclusion, we have used the AMI LA to observe a sample of five Lynds Dark Nebulae selected as candidates for spinning dust emission from the AMI SA sample of Lynds Dark Nebulae (Paper I). Towards two of these clouds (L1103 and L1111) we detect only patchy diffuse emission characteristic of the presence of a larger structure which has been mostly resolved out. Towards L675 we have observed flat spectrum compact cm-wave emission coincident with the SCUBA 850\\,$\\mu$m emission from the same region. These characteristics suggest that this source is associated with a stellar wind from a deeply embedded young protostar. We detect extended cm-wave emission to the North of the L944 SMM-1 protostar which displays spectral behaviour consistent with either spinning dust, or alternatively a collection of ultracompact {\\sc Hii} regions. L1246 shows an arc of cm-wave emission which is coincident with emission seen in \\emph{Spitzer} Band 4. We suggest that this is an example of emission from a population of PAH molecules, seen in emission lines in the \\emph{Spitzer} data, and emission as a consequence of rapid rotation of the molecules in the cm-wave data." }, "0910/0910.1919_arXiv.txt": { "abstract": "Spectroscopic long-slit observations of the dwarf Irr~galaxy IC~10 were conducted at the 6-m Special Astrophysical Observatory telescope with the SCORPIO focal reducer. The ionized-gas emission spectra in the regions of intense current star formation were obtained for a large number of regions in IC~10. The relative abundances of oxygen, N$^+$, and S$^+$ in about twenty HII~regions and in the synchrotron superbubble were estimated. We found that the galaxy-averaged oxygen abundance is $12+\\log(\\textrm{O/H}) = 8.17 \\pm 0.35$ and the metallicity is $Z=0.18 \\pm 0.14 Z_{\\odot} $. Our abundances estimated from the strong emission lines are found to be more reliable than those obtained by comparing diagnostic diagrams with photoionization models. ", "introduction": " ", "conclusions": "" }, "0910/0910.1634_arXiv.txt": { "abstract": "The accretion process onto spinning objects in Kerr spacetimes is studied with numerical simulations. Our results show that accretion onto compact objects with Kerr parameter (characterizing the spin) $|a| < M$ and $|a| > M$ is very different. In the super-spinning case, for $|a|$ moderately larger than $M$, the accretion onto the central object is extremely suppressed due to a repulsive force at short distance. The accreting matter cannot reach the central object, but instead is accumulated around it, forming a high density cloud that continues to grow. The radiation emitted in the accretion process will be harder and more intense than the one coming from standard black holes; e.g. $\\gamma$-rays could be produced as seen in some observations. Gravitational collapse of this cloud might even give rise to violent bursts. As $|a|$ increases, a larger amount of accreting matter reaches the central object and the growth of the cloud becomes less efficient. Our simulations find that a quasi-steady state of the accretion process exists for $|a|/M \\gtrsim 1.4$, independently of the mass accretion rate at large radii. For such high values of the Kerr parameter, the accreting matter forms a thin disk at very small radii. We provide some analytical arguments to strengthen the numerical results; in particular, we estimate the radius where the gravitational force changes from attractive to repulsive and the critical value $|a|/M \\approx 1.4$ separating the two qualitatively different regimes of accretion. We briefly discuss the observational signatures which could be used to look for such exotic objects in the Galaxy and/or in the Universe. ", "introduction": "It is widely believed that the final product of the gravitational collapse of matter is a black hole (BH). In classical general relativity (GR), astrophysical BHs should be completely characterized by just three parameters: the mass $M$, the charge $Q$, and the spin $J$. In this paper we focus on chargeless BHs. The spin is often replaced by the Kerr parameter $a = J/M$. In classical GR, the values of $M$ and $a$ cannot be completely arbitrary, as they must satisfy the relation $|a| < M$, which is the condition for the existence of the horizon. To see this we can examine the 3+1 dimensional Kerr solution. The position of the horizon is given by the expression~\\cite{mtw,lppt} \\be r_{H} = M + \\sqrt{M^2 - a^2} \\, . \\label{r-hor} \\ee It is clear that in (3+1)D spacetime the horizon cannot be formed if \\be M < |a| \\, . \\label{mass-limit} \\ee In the absence of a horizon, there would be naked singularities which are not allowed in GR. Indeed, if condition~(\\ref{mass-limit}) is fulfilled, the Kerr metric makes it possible to reach the physical singularity at $r=0$ from some large $r$ in finite time without crossing any horizon. One could thus consider closed time-like curves and violate causality (see e.g. section~66c of~\\cite{chandra} or ref.~\\cite{carter}). For this reason, usually some kind of cosmic censorship is assumed and naked singularities are forbidden~\\cite{penrose}. In particular, it is believed that naked singularities cannot be created by any physical process and therefore that the end-state of the gravitational collapse of matter is a Kerr BH with $|a| < M$~\\cite{penrose}. However, in this paper we consider objects which {\\it do} violate the Kerr bound, i.e. with $|a|>M$. We call them ``super-spinars'', as proposed in~\\cite{horava}: since they have no event horizon, by the standard definition they are not BHs. Our main motivation is simple. The singularity can be viewed as the place where new physics should be expected: here observer-independent quantities like the scalar curvature diverge, while GR presumably breaks down above the Planck scale. It is therefore not unreasonable to expect that causality is conserved, not because the collapsing matter can form only objects with $|a| < M$, but because actually the central singularity is replaced by some high curvature region due to some quantum gravity effects, see e.g.~\\cite{hn04,horava}. In this case, there is apparently no reason to believe that the final product of the gravitational collapse of matter cannot have $|a| > M$. Another possibility is that the collapsing matter forms a super-compact star with $|a| > M$ and exotic equations of state: now there is no central singularity, since the Kerr metric is a solution of Einstein equations only in vacuum; matter could have very exotic equation of state once it reaches densities so high that our knowledge of physics becomes inadequate. Actually, in general the metric at very small radii may deviate from the Kerr solution (the uniqueness theorem does not hold in absence of a regular horizon~\\cite{robinson}), but in our discussion we will neglect such a possibility. In this paper, we extend the studies started in~\\cite{bf09, bft09}. The goal is to examine the main differences between the cases $|a| < M$ and $|a| > M$. Previous papers~\\cite{bf09, bft09} discussed implications on the apparent shape. There it was found that, even if the bound is violated by a small amount, the shadow cast by the super-spinar (i.e. how it blocks light coming to us from an object behind it) changes significantly from the case with $|a| < M$: the shadow for the super-spinar is about an order of magnitude smaller as well as distorted. This distinction can be used as an observational signature in the search for these objects. Based on recent observations at mm wavelength of the super-massive BH candidate at the Center of the Galaxy~\\cite{doeleman}, the authors speculated on the possibility that it might violate the Kerr bound. In this paper we discuss the process of accretion in a Kerr background with arbitrary value of the Kerr parameter. For $|a|$ moderately larger than $M$, we find that the accreting matter cannot reach the central object, but is accumulated around it, forming a high density cloud. That may have interesting observational consequences. First, because of the high density and the high temperature of the plasma, the radiation produced in the accretion process can be much harder and more intense than the one coming from BHs. Second, there might be violent phenomena like bursts, when the amount of accumulated gas is large enough to gravitationally collapse. For higher values of $|a|$, the cloud evolves into a sort of disk, which is however very different from the usual disk of accretion around a BH: here the disk extends from $r \\approx M$ to the center, leading to rapid accretion, increasing efficiently the mass of the central object, and producing hard radiation at very small radii. Unlike the case of Kerr BHs, we do not know if super-spinars are stable under small perturbations. Previous work has found that some very rapidly rotating objects in { 3+1} and higher dimensions can be unstable~\\cite{emparan, cardoso}. To address this point regarding the super-spinars studied in this paper, one should do a linearized analysis of the perturbations of these objects. However, the conclusion would be determined by the boundary conditions at the surface of super-spinars, which presumably depend on the quantum theory of gravity and are therefore unknown. Such a question cannot thus be addressed at present: here we just assume that super-spinars are stable and we study the accretion process onto these objects. {\\it Conventions}: We use natural units $G_N = c = k_B = 1$. The metric has signature $(-+++)$. ", "conclusions": "In this paper we have studied the accretion process onto a spinning object. In particular, we have considered a Kerr spacetime with absolute value of the Kerr parameter $a$ (ratio of spin to mass) either smaller or larger than $M$, the mass of the object. In the first case, the spacetime contains a BH, in the second one a naked singularity. Our main motivation for considering the possibility of a naked singularity is based on the observation that the singularity is more likely a pathology of classical GR and that in the full theory it must be replaced by something else. We do not know how the central singularity is resolved, but our results are probably not significantly affected by the details of the correct theory. The only relevant quantity astrophysically is likely to be $|a|/M$. We can distinguish three cases: \\begin{enumerate} \\item BHs with $|a|/M<1$. We find the usual accretion picture: injected matter always reaches a quasi-steady state configuration, in which matter is lost behind the event horizon at the same rate as it enters into the computational domain. \\item Super-spinars with $1<|a|/M<1.4$. Here the gas cannot reach the central object, because of a repulsive force in the neighborhood of the center. As a result, the gas is efficiently accumulated around the super-spinar. That leads to the formation and growth of a high density cloud. However, the accumulation process will stop at some point. One possibility is that it is interrupted by violent events due to the gravitational collapse of the cloud onto the object. This could be associated with the formation of a new object, either a BH or a heavier super-spinar. Another possibility is that the accumulated matter creates an event horizon, hiding the object from the outside, and with no abundant release of energy. \\item Super-spinars with $|a|/M \\gtrsim 1.4$. Now the repulsive force around the center is no longer capable of preventing a regular accretion of the object. Our simulations find that the flow forms a high density thin disk on the equatorial plane and reaches a quasi-steady state, i.e. matter enters and leaves the computational domain at the same rate. This disk is much closer to the center of the object than in the case of a standard BH. \\end{enumerate}" }, "0910/0910.3574.txt": { "abstract": "We present a catalogue of variable stars in the near-infrared wavelength detected with overlapping regions of the 2MASS public images, and discuss their properties. The investigated region is in the direction of the Galactic center ($-30^\\circ \\lesssim l \\lesssim 20^\\circ, |b| \\lesssim 20^\\circ$), which covers the entire bulge. We have detected 136 variable stars, of which 6 are already-known and 118 are distributed in $|b| \\leq 5^\\circ$ region. Additionally, 84 variable stars have optical counterparts in DSS images. The three diagrams (colour-magnitude, light variance and colour-colour diagrams) indicate that most of the detected variable stars should be large-amplitude and long-period variables such as Mira variables or OH/IR stars. The number density distribution of the detected variable stars implies that they trace the bar structure of the Galactic bulge. ", "introduction": "The variable stars are useful tracers to study properties of the Galactic bulge. Among variables, the long-period variable stars (such as Semi-Regular Variables (SRVs), Miras and OH/IR stars) are detectable even in highly obscured region of the bulge owing to their high luminosities. These variable stars have been often used for investigating the Galactic bulge :e.g., measurement of the distance to the Galactic centre using Mira variables \\citep{Glass1995-MNRAS,Catchpole1999-IAUS,Groenewegen2005-AA,Matsunaga2009-pre}, a study of population of the bulge \\citep{Glass2001-MNRAS,Groenewegen2005-AA}. Therefore, a discovery of variable stars leads to provide us with important information about both variable star itself and the Galactic bulge. Due to a large amount of dust in the bulge, past searches for variable stars have been mainly focused on relatively unobscured fields such as Baade's Windows \\citep[e.g., ][]{Evans1976-MNRAS,Glass1995-MNRAS}. Near-infrared observation is also better for collecting variable stars in the inner bulge, since near-infrared light suffers relatively small interstellar absorption. \\citet{Glass2001-MNRAS} detected large-amplitude variable stars in a $24 \\times 24$ arcmin$^2$ area of the Galactic centre from K-band survey, and obtained $\\sim 400$ objects with periods and amplitudes. \\citet{Schultheis2000-AA} looked for variable stars by comparing only twice DENIS observations, and presented two catalogues with the property of their variable star candidates. However, past surveys have only investigated small restricted area (up to a few deg$^2$). Various imaging surveys have been performed in this two decades. In the imaging surveys, overlapping regions are produced for observing the sky without spatial gaps. In other words, there are regions observed multiple times in different epochs even if an imaging survey was performed only one time. Therefore, variable objects are expected to be detected with the overlapping regions. As claimed by \\citet{Schultheis2000-AA}, $\\sim 40\\%$ of variable stars can be recovered based on only twice-epoch magnitudes. Accordingly, the overlapping regions can be useful tools to search for variable stars. However, they have not been used in order to search for variable stars. When we can establish the search method based on overlapping region, it is also possible to search for variable stars using other survey data because survey images always contain overlapping regions. Furthermore, we can easily collect many variable stars in widely region (over a few tens of deg$^2$) with smaller time though we can extract less information compared with variable stars derived by an ordinary search (e.g., period, amplitude). In this paper, we present a catalogue of variable stars obtained from overlapping region of 2MASS public images. In Sect. \\ref{Data}, we introduce 2MASS observation and data, and discuss the photometric accuracy in an overlapping region. In Sect. \\ref{Data-Analysis}, after introducing the detection procedure and criteria, we show a result of search and estimate the interstellar extinctions. In Sect. \\ref{Probability}, we discuss detection probability based on twice-epoch magnitudes. In Sect. \\ref{Discussion}, we discuss the near-infrared properties of the detected variable stars and the spatial distribution in the bulge. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% ", "conclusions": "We have discovered variable stars using overlapping regions in the 2MASS public images. As a result of the investigation toward the Galactic centre ($-30^\\circ \\lesssim l \\lesssim 20^\\circ, |b| \\lesssim 20^\\circ$), we have detected 136 variable stars. Among which, 6 variable stars are already-known and 84 are accompanied by an optical counterpart in DSS images. The optical counterparts, however, are too faint to measure magnitudes (i.e., their brightness are nearby limiting magnitude of DSS images). The interstellar extinctions are estimated on the basis of the position of the upper giant branch, which are consistent with values of \\citet{Dutra2003-MNRAS}. Photometric properties of the detected variable stars indicate that most of them are large-amplitude AGB variables such as Mira variables and OH/IR stars in the Galactic bulge. Additionally, the bar-like structure of the bulge is detected by the number density distribution of the detected variable stars. This paper demonstrates that the search for variable stars using overlapping regions is an useful method. In this paper, there is a small number of detection due to the strict detection criteria, but it is possible to detect a lot of variable stars if we can judge the variability more precisely with tender criteria. Establishing this search method, we can discover variable stars in widely region. If extracting more variable stars in the bulge, the structure of the bulge might be revealed using them as tracers, since the detected variable stars in this paper trace the structure of the bulge." }, "0910/0910.1544_arXiv.txt": { "abstract": "{We present observations of two new single-lined eclipsing binaries, both consisting of an Am star and an M-dwarf, discovered by the Wide Angle Search for Planets transit photometry survey. Using WASP photometry and spectroscopic measurements we find that HD186753B has an orbital period of $P=1.9194$ days, a mass of $M=0.24\\pm0.02 M_{\\odot}$ and radius of $R=0.31^{+0.06}_{-0.06} R_{\\odot}$; and that TCY7096-222-1B has an orbital period of $P=8.9582$ days, a mass of between 0.29 and 0.54 $M_{\\odot}$ depending on eccentricity and radius of $R=0.263^{+0.02}_{-0.07} R_{\\odot}$. We find that the Am stars have relatively low rotational velocities that closely match the orbital velocities of the M-dwarfs, suggesting that they have been ``spun-down'' by the M-dwarfs.} ", "introduction": "The radius of a star is one of its most fundamental properties, yet for sub-solar masses models have not been able to provide accurate radius predictions. Citing the discrepancies between model and empirical radius measurements, Chabrier et al. (\\cite{chabrier}) found that large surface spot coverage decreases the photospheric temperature. The star compensates by increasing its radius to conserve radiative pressure. This was confirmed by an empirical activity-radius study by Lopez-Morales (\\cite{lopez-morales}). In addition, Berger et al. (\\cite{berger}) found a correlation between an increase in metalicity and a larger-than-expected radius. Because of their low intrinsic brightness, low-mass stars (LMS) are particularly difficult to study. LMS in eclipsing binary systems (EBLM), however, provide a direct way to obtain radius measurements and are therefore a valuable tool for testing models of stellar structure in the low-mass region. A by-product of wide-angle transit photometry planet-searching projects is the discovery of EBLMs (e.g. Fernandez et al. \\cite{newtrespaper}). The metallic-line Am stars are a class of peculiar A-type stars that are slow rotating, thought to have had their rotational velocity reduced by a near stellar companion. Spectroscopic orbits of many Am stars have been reported (e.g. Carquillat \\& Prieur \\cite{ampaper}, Renson \\& Manfroid \\cite{renson}) and LMS are thought to be responsible for reducing the rotational velocity of Am stars (Carquillat \\& Prieur \\cite{ampaper}). Here we report the discovery of two single lined A-M binaries, HD186753 and TYC7096-222-1, the first EBLMs discovered from the Wide Angle Search for Planets (WASP) planet-hunting project. There are four likely eclipsing A-M systems that have previously been announced; three of these systems were announced in Dreizler et al. (2002) and one in Pont et al. (2005). ", "conclusions": "\\label{discussion} The rotational velocity of A8 stars is expected to be $\\sim~200\\rm{km\\ s^{-1}}$ (Gray \\cite{gray}). The relatively low rotational velocities of HD186753A and TYC7096-222-1A, $v\\sin{i}=65.0\\pm5.0$ and $35.0\\pm5.0 \\rm{km\\ s^{-1}}$, respectively, suggests that they have been ``spun-down'' by their M-dwarf companions. The stellar rotational angular velocity of HD186753A is $(3.67\\pm0.65)\\times10^{-4} \\rm{rad\\ s^{-1}}$ which is greater than the M-dwarf orbital angular velocity of $(3.79\\pm0.01)\\times10^{-5} \\rm{rad\\ s^{-1}}$. The rotational angular velocity of TYC7096-222-1A is $(3.16\\pm0.47)\\times10^{-5} \\rm{rad\\ s^{-1}}$ which is also greater than the M-dwarf orbital angular velocity of $(8.12\\pm0.02)\\times10^{-6} \\rm{rad\\ s^{-1}}$. The synchronisation timescales are 2.8 Myr for HD186753 and 0.92 Gyr for TYC7096-222-1 (Zahn 1977), suggesting that each system is younger than the synchronisation time. Values of $\\log{g}$ can be used as an age estimate, but our value is not reliable enough. Higher SNR spectra is required to determine the age of the systems. The eccentricity of HD186753B, $e=0.269\\pm0.087$, is quite high for a short period binary and could also indicate a young stellar age as the circulization time is 95 Myr (Zahn 1977). The circulization time for TYC7096-222-1B is 295 Gyr. If TYC7096-222-1B has a mass of $M_{2}=0.54\\pm0.06 M_{\\odot}$ (using the $e=0.75$ model) then we would expect a secondary eclipse of $0.003\\pm1\\times{}10^{-4}$ mag deep. There appears to be no sign of this in the WASP photometry at any phase. TYC7096-222-1B therefore has a mass between $0.29-0.54 M_{\\odot}$ depending on the eccentricity. The certainty of $e$ for both objects could be improved with greater radial velocity phase coverage. The value of $e$ for HD186753B could be explained by a tertiary component either in the system or by a recent near-miss, although we find no evidence in our data for either scenario. Both primaries show supersolar Fe abundances, an underabundance of Ca and Sc and overabundances of Y and Ba. These abundances are typical of Am stars (e.g. Wolff \\cite{wolff}; Hundt \\cite{hundt}). Carquillat \\& Prieur (\\cite{ampaper}) found that the mean companion mass to an Am spectral type is $0.8\\pm{0.5}M_{\\odot}$ with a mean orbital period of 1.995 days. By these stellar properties HD186753 and TYC7096-222-1 are fairly typical Am systems. The radii of stars with masses below $0.3-0.35\\rm{M_{\\odot}}$ agree well with the Baraffe et al. (\\cite{baraffe}) isochrone models (Lopez-Morales \\cite{lopez-morales}). The mass-radius relation TYC7096-222-1B agree within $1\\sigma$ of the Baraffe et al. (\\cite{baraffe}) isochrones, as shown in Fig. \\ref{m-r}. Whilst the mass-radius relation of HD186753B, however, is only just within $1\\sigma$ of the Baraffe et al. (\\cite{baraffe}) isochrones. M-dwarfs in binaries are found to be more active than solitary M-dwarfs (Lopez-Morales \\cite{lopez-morales}). Increased activity causes a decrease in photospheric temperature, which then causes an increase in radius to conserve radiative pressure (Ribas \\cite{Ribas}). X-ray activity is an indicator of stellar activity, although no X-ray data has been published on either object. Being in a tight orbit, the photosphere of HD186753B is strongly irradiated by HD186753A. The radiation intensity received from HD186753A at the photosphere of HD186753B, and also therefore the radiative pressure, is $1.32\\pm0.02$ times the radiation generated; the value for TYC7096-222-1 is lower at $0.72\\pm0.04$. Higher precision radial velocity and photometry than the values given here are required to ascertain as to whether this extra source of heating, or increased levels of activity are affecting the mass-radius relation of the M-dwarfs. \\begin{figure} \\centering \\resizebox{\\columnwidth}{!}{\\includegraphics[angle=270,bb=50 50 554 770]{M_vs_R.ps}} \\caption{Mass-radius relationship for solitary stars, EBLMs, HD186753B and TCY7096-222-1 superimposed on 5 Gyr, [M/H]=0, $L_{\\rm{mix}}=1H_{p}$ Baraffe et al. (\\cite{baraffe}) isochrones. The thick black line indicates the range of masses for TYC7096-222-1B depending on the eccentricity. M-dwarfs with uncertainties in mass $>0.03 M_{\\odot}$ have been omitted from the figure for illustrative clarity.} \\label{m-r} \\end{figure} \\begin{table*} \\begin{tabular}{llll} Parameter & HD186753B & TYC7096-222-1B ($e=0.00$) & TYC7096-222-1B ($e=0.75$) \\\\ \\hline\\hline Eclipse epoch (HJD) & $3940.40144^{+0.00091}_{-0.00081}$ & $4373.01637^{+0.00066}_{-0.00087}$ & - \\\\ Orbital period (days) & $1.9193851^{+0.0000412}_{-0.0000393}$ & $8.9582591^{+0.0000346}_{-0.0000314}$ & - \\\\ Eclipse duration (days) & $0.1662^{+0.0038}_{-0.0031}$ & $0.2430^{+0.0039}_{-0.0042}$ & - \\\\ Secondary/primary area ratio, $(R_{2}/R_{1})^2$ & $0.0148^{+0.0005}_{-0.0003}$ & $0.0251^{+0.0006}_{-0.0005}$ & - \\\\ Impact parameter, $b (R_{*})$ & $0.264^{+0.191}_{-0.154}$ & $0.250^{+0.157}_{-0.142}$ & - \\\\ Stellar reflex vel., $\\rm{K_{1}}$ ($\\rm{km\\ s^{-1}}$) & $-27.449\\pm1.751$ & $-20.341\\pm2.974$ & $-58.432\\pm6.122$ \\\\ Centre-of-mass vel., $\\gamma$ ($\\rm{km\\ s^{-1}}$) & $-14.641\\pm2.980$ & $4.074\\pm1.533$ & $8.776\\pm0.748$ \\\\ Orbital separation, $a$ (AU) & $0.0370^{+0.0012}_{-0.0013}$ & $0.0990^{+0.0031}_{-0.0034}$ & - \\\\ Orbital inclination, $i$ (deg) & $87.09^{+1.69}_{-1.79}$ & $89.00^{+0.57}_{-0.73}$ & - \\\\ Orbital eccentricity, $e$ & $0.269\\pm0.087$ & (fixed=0.00) & (fixed=0.75) \\\\ Longitude of periastron, $\\omega$ (deg) & $166.7\\pm5.9$ & (not fitted) & $167.2\\pm61$ \\\\ Stellar Mass, $M_{2}(M_{\\odot})$ & $0.236\\pm0.017$ & $0.286\\pm0.019$ & $0.544\\pm0.057$ \\\\ Stellar Radius, $R_{2}(R_{\\odot})$ & $0.307^{+0.057}_{-0.057}$ & $0.263^{+0.020}_{-0.071}$ & - \\\\ Luminosity ratio, $L_{1}/L_{2}$ & $979.7$ & $520.4$ & $361.9$ \\\\ \\end{tabular} \\caption{\\label{paramtable} Parameters of HD186753B and TYC7096-222-1B and their orbits. The parameters of TYC7096-222-1B determined from the photometry are the same for both eccentricities.} \\end{table*}" }, "0910/0910.4966_arXiv.txt": { "abstract": "We utilize color information for an HI-selected sample of 195 galaxies to explore the star formation histories and physical conditions that produce the observed colors. We show that the HI selection creates a significant offset towards bluer colors that can be explained by enhanced recent bursts of star formation. There is also no obvious color bimodality, because the HI selection restricts the sample to bluer, actively star forming systems, diminishing the importance of the red sequence. Rising star formation rates are still required to explain the colors of galaxies bluer than $g-r <$ 0.3. We also demonstrate that the colors of the bluest galaxies in our sample are dominated by emission lines and that stellar population synthesis models alone (without emission lines) are not adequate for reproducing many of the galaxy colors. These emission lines produce large changes in the $r-i$ colors but leave the $g-r$ color largely unchanged. In addition, we find an increase in the dispersion of galaxy colors at low masses that may be the result of a change in the star formation process in low-mass galaxies. ", "introduction": "The star formation history, metallicity, and current star formation rate all contribute to the observed colors of a galaxy. Remarkably, these factors work in concert to yield a fairly well defined locus in galaxy color-color space (Strateva et al.~2001). Deviations from this locus as well as the morphology of the locus itself, can lead to significant insight into the underlying processes taking place within galaxies. Many previous studies have utilized the broad-band colors of galaxies to investigate the star formation histories and metallicities of galaxies (e.g. Tinsley 1972; Searle et al.~1973; Tinsley \\& Gunn 1976; Balcells \\& Peletier 1994; Roberts \\& Haynes 1994; de Jong 1996; Bell \\& de Jong 2000; Galaz et al.~2002; Gavazzi et al.~2002; Bell et al.~2003; MacArthur et al.~2004; Zackrisson, Bergvall \\& Ostlin 2005; Driver et al.~2006; Skibba et al.~2008). The advent of large surveys such as the Sloan Digital Sky Survey (SDSS; York et al.~2000) and the Two Micron All Sky Survey (2MASS; Skrutskie et al.~2006) has created a wealth of uniform data and the statistical foothold to investigate the bimodality of galaxies (Baldry et al.~2004; Kauffmann et al.~2003, 2004), luminosity function (Blanton et al.~2001; Ball et al.~2006), and the average properties of nearby galaxies (Blanton et al.~2003a, 2003b, 2005; Geha et al.~2006; Maller et al.~2008; Skibba et al.~2008). While the large optical and infrared surveys have made large contributions to our understanding of galaxy evolution, they only trace the stellar component of galaxies and do not trace other baryonic material such as cold gas. Galaxies in the local universe span a range of star formation histories -- from blue, gas-rich, low-surface-brightness (LSB) galaxies that are slowly turning their gas into stars with low star forming efficiencies, to red, gas-poor galaxies, that have formed the bulk of their stars in the past. Stars dominate the visible light output of most galaxies, and thus galaxies detected by traditional optical or infrared imaging have well developed stellar populations. In contrast, the natural way to identify gas-rich, less evolved galaxies is by their 21 cm HI radio emission. Aside from its importance for global star formation, a sample of galaxies with both gaseous and stellar information allows for a more complete census of the local baryons (a constraint vital to the calibration of n-body simulations; Governato et al.~2007; Brooks et al.~2009). Previous studies have combined large HI surveys with optical and infrared samples, namely the Arecibo Duel Beam and Slice Surveys with the Two Micron All Sky Survey (2MASS; Jarrett et al.~2000; Rosenberg et al.~2005) and the merging of HIPASS with SuperCOSMOS (Hambly et al.~2001; Hambly, Irwin \\& MacGillivary 2001b; Hambly et al.~2001c; Doyle et al.~2005). Rosenberg et al.~(2005) were able to probe the baryonic content of a large sample of galaxies, but were limited by the shallow depth of 2MASS, which does not have data for many of the LSB galaxies in the sample. The HIPASS/SuperCOSMOS sample of Doyle et al.~(2005) contains optical data for more than 3600 HI selected galaxies but also suffers from the shallow depth of the SuperCOSMOS optical data. Recent studies have combined the Parkes HI Equatorial Survey (ES) with the SDSS (Disney et al.~2008; Garcia-Appadoo et al.~2009; West et al.~2009; hereafter W09). The deep SDSS optical data provides information about the stellar content for \\emph{all} ES HI sources where the two surveys overlap. In addition, the uniform, accurate, and well-calibrated photometry of the HI-selected ES/SDSS sample allows for a more detailed exploration of the factors affecting galaxy colors, particularly for galaxies with large reservoirs of gas. The optical colors are reasonably sensitive to age, although IR colors are needed to constrain metallicity (Bell et al.~2003). There is thus an unavoidable degeneracy between age and metallicity when using only the optical colors available with SDSS (Bell \\& de Jong 2001; Bell et al.~2003). Our sample does not have a complete set of near IR counterpart data and some of our results will reflect this limitation. In this paper, we briefly describe the ES/SDSS sample in \\S2 and examine the colors of HI-selected galaxies by comparing them to stellar population synthesis models (\\S3.1) and the colors of optically selected galaxies (\\S3.2). We also investigate how line emission affects the broadband colors of gas-rich galaxies (\\S3.3). We demonstrate an increased dispersion in the colors of galaxies at low masses, and investigate its possible origins (\\S3.4). We summarize and discuss our results in \\S4. ", "conclusions": "We used the population synthesis models of Bruzual \\& Charlot (2003) to model the SFHs and metallicities of galaxies in the ES/SDSS sample. We found that red galaxies have super-solar metallicities and have SFHs that are consistent with them forming the bulk of their stars in the distant past. These red galaxies have presumably exhausted their early gas supply but have recently acquired gas through mergers and infall. Although the infalling gas is likely low metallicity and could in principle dilute the metallicity, the small amount of gas in red galaxies (low gas fractions) is not enough to make an observable difference in the average metallicity of the ES/SDSS systems. Bluer galaxies have lower metallicities and their mean stellar ages are younger (increasing $\\tau$ values). The gas fractions suggests that these blue systems have at least as much gas as they do stars. Previous studies have indicated that these types of galaxies are less massive and have disks that are stable against collapse, making their star formation much less efficient. They have also likely retained much of their initial HI and are slowly converting it into stars as is suggested by the large (or negative) $\\tau$ values. As galaxies pass the threshold for disk stability, their star formation becomes sporadic and there is a clear dispersion in the colors due to variations in burst age, burst strength and superimposed emission lines. The idealistic Bruzual and Charlot models are likely only loose guides for these systems, as their star formation histories no longer follow smooth exponential functions. Figure \\ref{bc1} indicates that there are a few galaxies in the regime of exponentially increasing SFR (negative $\\tau$ values). This is not surprising as there are few mechanisms that will increase the SFR of a galaxy. The likely culprit for these systems is the infall of gas. As gas is accreted, the gas densities in the galaxies will increase and local gravitational collapse will become more efficient. This will increase the SFR as well as the eventual metallicity of the galaxy, pushing its $g-r$ colors to the blue (more recent star formation) and its $r-i$ colors slightly to the red (more metals). We also showed that HI selected galaxies are offset from the SDSS galaxy locus, especially at the red end, and that this is likely due to bursts of star formation in the past few hundred Myr. The gas that induces these recent bursts is not primordial and is best explained by the accretion of gas rich dwarfs. The bluest galaxies in the ES/SDSS sample are not explained by population synthesis models alone. Their colors can be modeled only with the inclusion of emission lines. We showed that emission line spectra with reasonable SFRs can explain the colors of the bluest galaxies in our sample. We also showed that the distribution of galaxies at the red end of the color-color locus has a very small dispersion that continues to rise into the blue regime. This change in dispersion appears to correlate with stellar mass, rotation velocity, surface brightness and especially stellar surface density. It is possible that the change in dispersion is evidence for a significant change in the way stars form in galaxies at a given mass scale. Massive galaxies have unstable disks and efficiently convert most of their primordial reservoirs of gas into stars in the first few Gyrs after their formation. These galaxies are bulge dominated and have redder, older and more massive populations of stars. We do see a sharp transition in the color distribution at a rotation velocity of 80 km~s$^{-1}$. This transition may be related to previous results that have found sharp transitions in galaxy properties as a function of rotational velocity (Dalcanton et al.~2004; Lee et al.~2007). However, further investigations are required to explore the various velocity transitions as a function of galaxy properties. It may also be the case that the increase in dispersion is nothing more than a selection effect related to the surface brightness. As the surface brightness of a system decreases, the effect of a single burst of stars on the color becomes increasingly large, possibly explaining the increase in scatter. This effect might explain the onset of emission line dominated colors in the bluest galaxies as they can influence of the colors of galaxies devoid of massive stellar populations. We note that the change in color dispersion is not easily seen in the volume selected SDSS data. Applying the HI selection identifies a low dispersion subsample of the SDSS galaxies, most notably at the red end. As mentioned above, the red HI selected galaxies appear to be bluer in $g-r$ than the ``main'' SDSS sample. These two features are likely related. It is possible that the large distribution in color at the red end of the SDSS main sample is due to the diversity of gas content. Because most of the red galaxies in SDSS exhausted their original supply of gas long ago, this dispersion may be an indication of the spread in time since the last major gas infall. We leave further discussion of the ``color offset'' to future study." }, "0910/0910.3893_arXiv.txt": { "abstract": "Sagittarius~A* is the source of near infrared, X-ray, radio, and (sub)millimeter emission associated with the supermassive black hole at the Galactic Center. In the submillimeter regime, Sgr~A* exhibits time-variable linear polarization on timescales corresponding to $< 10$ Schwarzschild radii of the presumed $4 \\times 10^6$~M$_\\sun$ black hole. In previous work, we demonstrated the potential for total-intensity (sub)millimeter-wavelength very long baseline interferometry (VLBI) to detect time-variable -- and periodic -- source structure changes in the Sgr~A* black hole system using nonimaging analyses. Here we extend this work to include full polarimetric VLBI observations. We simulate full-polarization (sub)millimeter VLBI data of Sgr~A* using a hot-spot model that is embedded within an accretion disk, with emphasis on nonimaging polarimetric data products that are robust against calibration errors. Although the source-integrated linear polarization fraction in the models is typically only a few percent, the linear polarization fraction on small angular scales can be much higher, enabling the detection of changes in the polarimetric structure of Sgr~A* on a wide variety of baselines. The shortest baselines track the source-integrated linear polarization fraction, while longer baselines are sensitive to polarization substructures that are beam-diluted by connected-element interferometry. The detection of periodic variability in source polarization should not be significantly affected even if instrumental polarization terms cannot be calibrated out. As more antennas are included in the (sub)mm-VLBI array, observations with full polarization will provide important new diagnostics to help disentangle intrinsic source polarization from Faraday rotation effects in the accretion and outflow region close to the black hole event horizon. ", "introduction": "The Galactic Center source Sagittarius~A* (Sgr~A*) provides the best case for high-resolution, detailed observations of the accretion and outflow region surrounding the event horizon of a black hole. There are several compelling reasons to observe Sgr~A* with very long baseline interferometry (VLBI) at (sub)millimeter\\footnote{We shall henceforth use the term ``millimeter'' to denote wavelengths of 1.3~mm of shorter (in contrast with observations at 3~mm and 7~mm, which are sometimes also referred to as ``millimeter'' wavelengths).} wavelengths. The spectrum of Sgr~A* peaks in the millimeter \\citep[][and references therein]{markoff07}. Interstellar scattering, which varies as the wavelength $\\lambda^2$, becomes less than the fringe spacing of the longest baseline available to VLBI in the millimeter-wavelength regime. Indeed, VLBI on the longest baselines available at 345~GHz probes scales of twice the Schwarzschild radius ($R_\\mathrm{S}$) for a $4 \\times 10^6$~M$_\\sun$ black hole. From previous observations at 230~GHz, it is known that there are structures on scales smaller than a few $R_\\mathrm{S}$ \\citep{doeleman08}. Such high angular resolution, presently unattainable by any other method (including facility instruments such as the Very Long Baseline Array), is necessary to match the expected spatial scales of the emitting plasma in the innermost regions surrounding the black hole and will be required to unambiguously determine the inflow/outflow morphology and permit tests of general relativity. This sensitivity to small spatial scales also makes millimeter polarimetric VLBI possible. Although the linear polarization fraction of Sgr~A* integrated over the entire source is only a few percent \\citep[e.g.,][]{marrone07}, the fractional polarization on small angular scales is likely much larger. In general, relativistic accretion flow models predict that the electric vector polarization angle (EVPA) will vary along the circumference of the accretion disk \\citep{bromley01,broderick05,broderick06}, indicating that single-dish observations and connected-element interferometers probably underestimate linear polarization fractions due to beam depolarization. Polarized synchrotron radiation coming from Sgr~A* was detected by \\citet{aitken00} at millimeter and submillimeter wavelengths. Multiple observations since then have demonstrated that the polarized emission is variable on timescales from hours to many days \\citep{bower05,macquart06,marrone06a,marrone07,marrone08}. In one case, the timescale of variability and the trace of polarization in the Stokes $(Q,U)$ plane of the millimeter-wavelength emission are suggestive of the detection of an orbit of a polarized blob of material \\citep{marrone06b}. Near infrared observations by \\citet{trippe07} are also consistent with a hot spot origin for periodic variability. It is possible that connected-element interferometry may suffice to demonstrate polarization periodicity, but millimeter-wavelength VLBI, which effectively acts as a spatial filter on scales of a few to a few hundred $R_\\mathrm{S}$, can be more sensitive to changing polarization structures. Initial millimeter VLBI observations of Sgr~A* will necessarily utilize non-imaging analysis techniques, for reasons outlined in \\citet[][henceforth Paper~I]{doeleman09}. One way to do this is to analyze so-called interferometric ``closure quantities,'' which are relatively immune to calibration errors \\citep{rogers74,rogers95}. In Paper~I, we considered prospects for detecting the periodicity signature of a hot spot orbiting the black hole in Sgr~A* via closure quantities in total-intensity millimeter-wavelength VLBI. In the single polarization case, it is necessary to construct closure quantities from at least three or four antennas in order to produce robust observables, since the timescales of atmospheric coherence and frequency standard stability do not permit standard nodding calibration techniques. Closure quantities can be used in full-polarization observations as well, but it is also possible to construct robust observables on a baseline of two antennas by taking visibility ratios between different correlation products. In this work, we extend our techniques to explore polarimetric signatures of a variable source structure in Sgr~A*, with emphasis on ratios of baseline visibilities. ", "conclusions": "(Sub)millimeter-wavelength VLBI polarimetry is a very valuable diagnostic of emission processes and dynamics near the event horizon of Sgr~A*. We summarize the findings in this paper as follows: \\begin{itemize} \\addtolength{\\itemsep}{-0.8ex} \\item Millimeter-wavelength polarimetric VLBI can detect changing source structures. Despite low polarization fractions seen with connected-element interferometry, the much higher angular resolution data provided by VLBI will be far less affected by beam depolarization and contamination from dust polarization. Polarimetric VLBI provides an orthogonal way to detect periodic structural changes as compared with total intensity VLBI. \\item Ratios of cross- to parallel-hand visibilities are robust baseline-based observables. Short VLBI baselines approximately trace the integrated polarization fraction and position angle of the inner accretion flow of Sgr~A*, while longer VLBI baselines resolve smaller structures. \\item Calibration of instrumental polarization terms is not necessary to detect a changing source structure, including periodicity, in Sgr~A*. \\item Polarimetric VLBI may be able to disentangle the effects of rotation measure from intrinsic source polarization. Initial results will likely come from observations of the timescale of polarimetric variability. If the initial array is expanded to allow high-fidelity imaging, polarimetric VLBI may be able to map the Faraday rotation region and directly infer the density and magnetic field structure of the emitting region in Sgr~A*. \\end{itemize}" }, "0910/0910.5292_arXiv.txt": { "abstract": "We obtained SDSS spectra for a set of 37 radio-quiet quasars (RQQSOs) that had been previously examined for rapid small scale optical variations, or microvariability. Their \\hbeta and \\mgii emission lines were carefully fit to determine line widths (FWHM) as well as equivalent widths (EW) due to the broad emission line components. The line widths were used to estimate black hole masses and Eddington ratios, $\\ell$. Both EW and FWHM are anticorrelated with $\\ell$. The EW distributions provide no evidence for the hypothesis that a weak jet component in the RQQSOs is responsible for their microvariability. ", "introduction": "Over the past 15 years there have been rather extensive examinations of a significant sample of radio-quiet QSOs (RQQSOs) and Seyfert galaxies for small brightness changes (typically 0.02 mag) over short times (a few hours) (e.g., Stalin et al.\\ 2004b; Carini et al.\\ 2007; Ram{\\'i}rez et al.\\ 2009). This phenomenon of microvariability, or intranight optical variability (INOV) was first confirmed for blazars (e.g., Miller, Carini \\& Goodrich 1989; Carini 1990) for which microvariability almost certainly arises from the relativistic jet (e.g., Marsher, Gear \\& Travis 1992), at least when the source is in an active state; however, in low states it is possible that rapid fluctuations are due to processes originating in or just above the accretion disc (for a review, see Wiita 2006). Since RQQSOs lack significant jets, the microvariability in radio-quiet objects may arise from processes on the accretion disc itself, and thus could possibly be used to probe the discs (e.g., Gopal-Krishna, Wiita \\& Altieri 1993; Stalin et al.\\ 2004a; Kelly, Bechtold \\& Siemiginowska 2009). Recently Carini et al.\\ (2007) reported new observations for several sources and also compiled a sample of 117 objects from the literature which have been searched for microvariability (their Table 3). Of these, 47 are classified as Seyfert galaxies, 6 as broad absorption line (BAL) QSOs, and 64 as QSOs. In addition, in order to learn which classes might be the most fruitful for making further searchs for microvariability they also classified their sample in terms of following: redshift distribution; radio loudness, through $R$, the ratio of the radio [5 GHz] flux to the optical [4400\\AA] flux); optical magnitude, $m_V$; luminosity; and observing strategy. In their entire sample 21.4\\% of the objects were found to exhibit microvariability, but among objects classified as Seyfert galaxies, BAL QSOs and QSOs, microvariability was found in 17\\%, 50\\% and 23.2\\%, respectively (Carini et al.\\ 2007). The observed high fraction of microvariations in BAL QSOs (although the sample is quite small) suggests that while planning microvariability studies to investigate the physical processes in or near the accretion disc, it might be worthwhile to invest more observing time on the BALQSO class. \\par With improvements in observation quality as well as in sample size, constraints on the models capable of producing the intranight variability have also improved over last decade. Recently, Czerny et al.\\ (2008) have used non-simultaneous optical and X-ray data of 10 RQQSOs with confirmed INOV to compare observational constraints on the variability properties with the predictions of theoretical models such as: (i) irradiation of an accretion disc by a variable X-ray flux (e.g., Rokaki, Collin-Souffrin \\& Magnan 1993; Gaskell 2006); (ii) an accretion disc instability (e.g., Mangalam \\& Wiita 1993); (iii) the presence of a weak Doppler boosted jet, or ``blazar component'' (e.g., Gopal-Krishna et al.\\ 2003). Their investigation suggests that a blazar component model yields the highest probability of detecting INOV. In this blazar component scenario, spectral properties of the sources can play a crucial role in constraining the models further. For instance, if blazar components are dominating the variability of RQQSOs, then, due to the increase in the continuum level, one would expect emission lines to be diluted. Therefore smaller equivalent widths (EWs) of prominent emission lines such as \\hbeta and \\mgii should be detected in sources that showed microvariability when compared to their average values in the whole sample. If we take the extreme case of BL Lacertae objects, which often lack observable emission lines and are usually defined as objects that have no emission line with an EW $\\geq$5\\AA \\ (e.g., Stickel et al.\\ 1991; cf., March{\\~a} et al.\\ 1996), this dilution by the jet component is understood to be severe. So it becomes very important to test whether microvariability of RQQSOs has any correlation with spectral parameters such as EW and full width at half maximum (FWHM) of prominent emission lines. The \\hbeta and \\mgii emission lines are very promising for such investigations, as these lines have also been found to be very useful in estimating other key parameters of AGN central engines such as black hole (BH) mass (e.g., McGill et al.\\ 2008) and Eddington ratio (e.g., Dong et al.\\ 2009a, 2009b). The average \\hbeta EW of a large sample of quasars is found to be around 62.4\\AA \\ (Foster et al.\\ 2001), so any correlation of \\hbeta EW or FWHM {\\bf would} not only give insight about the nature of variability, such as the presence or absence of blazar components, but also {\\bf would} be very useful for making a promising sample for future microvariability studies. Similarly the measurements of BH masses for RQQSOs that show microvariability may well be another important constraint {\\bf on} the models trying to understand the nature of their optical microvariability. Here we have worked toward these goals by exploiting the optical spectra available from Sloan Digital Sky Survey (SDSS) Data Release 7 (DR7; Abazajian et al.\\ 2009) with careful spectral modeling of the \\hbeta and \\mgii emission line regions. First we aim to investigate any effect of these key spectral parameters (e.g., EW and FWHM) on microvariability of RQQSOs. Second we estimate other relevant AGN parameters such as the black hole mass, M$_{bh}$, and the Eddington ratio in the context of our RQQSOs sample, which has been extensively searched for microvariability. \\par The paper is organized as follows. Section 2 describes the data sample and selection criteria while Section 3 gives details of our spectral fitting procedure. In Section 4 we focus on BH mass measurements and in Section 5 we give estimates of Eddington ratios and of BH growth times. Section 6 gives a discussion and conclusions. Throughout, we have used a cosmology with $H_{\\rm 0}$=70 km\\,s$^{-1}$\\,Mpc$^{-1}$, $\\Omega_{\\rm M}$=0.3 and $\\Omega_{\\rm \\Lambda}$=0.7. \\begin{table*} \\centering \\begin{minipage}{140mm} \\caption{ Our sample of radio-quiet QSOs and Seyfert galaxies from the compilation of Carini et al.\\ (2007)} \\label{lab:tabsamp} \\begin{tabular}{@{}llllllllllcl@{}} \\hline \\multicolumn{1}{c}{QSO Name\\footnote{Name from V{\\'e}ron-Cetty \\& V{\\'e}ron (2006)}} & {$z_{em}$} & \\multicolumn{3}{c}{$\\alpha_{2000}$} & \\multicolumn{3}{c}{$\\delta_{2000}$} &{m$_{V}$} & {M$_{V}$} & {Variable?} & {Class} \\\\ \\hline \\\\ Mrk 1014 & 0.163 & 01& 59& 50.2 & $+$00& 23& 41& 15.9 & $-$23.8 & Y& Sy 1 \\\\ US 3150 & 0.467 & 02& 46& 51.9 & $-$00& 59& 31& 17.1 & $-$25.0 & N& NLS1 \\\\ US 3472 & 0.532 & 02& 59& 37.5 & $+$00& 37& 36& 16.8 & $-$25.7 & N& QSO \\\\ Q J0751$+$2919 & 0.912 & 07& 51& 12.3 & $+$29& 19& 38& 16.2 & $-$27.7 & Y& QSO \\\\ PG 0832$+$251 & 0.331 & 08& 35& 35.9 & $+$24& 59& 41& 16.1 & $-$25.5 & Y& QSO \\\\ US 1420 & 1.473 & 08& 39& 35.1 & $+$44& 08& 11& 17.5 & $-$27.3 & N& QSO \\\\ US 1443 & 1.564 & 08& 40& 30.0 & $+$46& 51& 13& 17.2 & $-$27.8 & N& QSO \\\\ US 1498 & 1.406 & 08& 42& 15.2 & $+$45& 25& 44& 17.7 & $-$26.7 & N& QSO \\\\ US 1867 & 0.513 & 08& 53& 34.2 & $+$43& 49& 01& 16.9 & $-$25.2 & N& Sy 1 \\\\ PG 0923$+$201 & 0.192 & 09& 25& 54.7 & $+$19& 54& 04& 15.5 & $-$24.8 & Y& Sy 1 \\\\ PG 0931$+$437 & 0.456 & 09& 35& 02.6 & $+$43& 31& 11& 16.0 & $-$26.4 & N& QSO \\\\ PG 0935$+$416 & 1.966 & 09& 38& 57.0 & $+$41& 28& 21& 16.8 & $-$28.8 & N& QSO \\\\ CSO 233 & 2.030 & 09& 39& 35.1 & $+$36& 40& 01& 18.4 & $-$27.3 & N& QSO \\\\ CSO 18 & 1.300 & 09& 46& 36.9 & $+$32& 39& 51& 17.0 & $-$27.9 & N& QSO \\\\ US 995 & 0.226 & 09& 48& 59.4 & $+$43& 35& 18& 16.9 & $-$23.4 & Y& QSO \\\\ PG 0946$+$301 & 1.220 & 09& 49& 41.1 & $+$29& 55& 19& 16.2 & $-$28.2 & N& BAL QSO \\\\ CSO 21 & 1.190 & 09& 50& 45.7 & $+$30& 25& 19& 17.3 & $-$27.3 & N& QSO \\\\ TON 34 & 1.925 & 10& 19& 56.6 & $+$27& 44& 02& 15.7 & $-$29.8 & Y& QSO \\\\ PG 1049$-$005 & 0.357 & 10& 51& 51.5 & $-$00& 51& 17& 15.8 & $-$25.7 & N& Sy 1 \\\\ TON 52 & 0.434 & 11& 04& 07.0 & $+$31& 41& 11& 17.3 & $-$24.9 & Y& QSO \\\\ PG 1206$+$459 & 1.155 & 12& 08& 58.0 & $+$45& 40& 36& 15.7 & $-$28.4 & N& QSO \\\\ UM 497 & 2.022 & 12& 25& 18.4 & $+$02& 06& 57& 17.7 & $-$28.0 & N& QSO \\\\ PG 1248$+$401 & 1.032 & 12& 50& 48.3 & $+$39& 51& 40& 16.3 & $-$27.6 & N& QSO \\\\ CSO 174 & 0.653 & 12& 51& 00.3 & $+$30& 25& 42& 17.0 & $-$26.1 & N& QSO \\\\ CSO 179 & 0.782 & 12& 53& 17.5 & $+$31& 05& 50& 17.0 & $-$26.4 & N& QSO \\\\ Q 1252$+$0200 & 0.345 & 12& 55& 19.7 & $+$01& 44& 12& 16.2 & $-$25.3 & Y& QSO \\\\ PG 1254$+$047 & 1.018 & 12& 56& 59.9 & $+$04& 27& 34& 16.3 & $-$27.6 & N& BAL QSO \\\\ PG 1259$+$593 & 0.472 & 13& 01& 12.9 & $+$59& 02& 07& 15.9 & $-$26.2 & N& QSO \\\\ PG 1307$+$085 & 0.154 & 13& 09& 47.0 & $+$08& 19& 49& 15.1 & $-$24.6 & N& Sy 1 \\\\ PG 1309$+$355 & 0.183 & 13& 12& 17.7 & $+$35& 15& 20& 15.6 & $-$24.6 & N& Sy 1.2 \\\\ CSO 879 & 0.549 & 13& 21& 15.8 & $+$28& 47& 20& 16.7 & $-$25.8 & N& QSO \\\\ PG 1338$+$416 & 1.204 & 13& 41& 00.8 & $+$41& 23& 14& 16.8 & $-$27.5 & N& QSO \\\\ CSO 448 & 0.316 & 14& 24& 55.6 & $+$42& 14& 05& 17.0 & $-$24.3 & Y& QSO \\\\ PG 1444$+$407 & 0.267 & 14& 46& 46.0 & $+$40& 35& 06& 16.1 & $-$24.7 & N& Sy 1 \\\\ PG 1522$+$101 & 1.328 & 15& 24& 24.5 & $+$09& 58& 30& 16.2 & $-$28.4 & N& QSO \\\\ Q 1628.5$+$3808 & 1.461 & 16& 30& 13.6 & $+$37& 58& 21& 17.7 & $-$27.2 & Y& QSO \\\\ PG 1630$+$377 & 1.478 & 16& 32& 01.1 & $+$37& 37& 49& 16.3 & $-$28.6 & N& QSO \\\\ \\hline \\end{tabular} \\end{minipage} \\end{table*} ", "conclusions": "Several recent papers have used the very large numbers of quasars discovered by modern surveys to estimate their BH masses using virial approaches (e.g., Shen et al.\\ 2008; Fine et al.\\ 2008; Vestergaard \\& Osmer 2009). The large numbers of quasars whose spectra are fit in these particular papers (between 1100 and almost 57,700) have allowed for very important new conclusions to be drawn about quasar demography. Shen et al.\\ (2008) have found that the line widths of \\hbeta and \\mgii follow lognormal distributions with very weak dependencies on redshift and luminosity. Their comparison of BH masses of radio-loud quasars (RLQSOs) and BALs with those of ``ordinary\" quasars (RQQSOs) shows that the mean of virial masses of RLQSOs is 0.12 dex larger than that of ordinary quasars, while that of BALs is indistinguishable from that of ordinary quasars. Fine et al.\\ (2008) used \\mgii lines to estimate BH masses, and found that the scatter in measured BH masses is luminosity dependent, showing less scatter for more luminous objects. Vestergaard \\& Osmer (2009) have studied the mass functions of BH masses at different redshifts, and found evidence for cosmic downsizing in their cosmic space density distributions. The sample we have considered is much smaller, but each member has been carefully selected to be among the special group of RQQSOs and Seyfert galaxies already examined for optical microvariability (e.g., Carini et al.\\ 2007). This criterion demands that modest aperture (usually 1--2 m) telescopes can make precise photometric measurements in just a few minutes, and so limits the members to the rare QSOs with bright apparent magnitudes (usually $m_V < 17.5$). In addition, special care was taken in the fitting of the line profiles as discussed in Sections 3.1 and 3.2. As the precision of CCD based differential photometry has typically improved to better than 0.01 mag for these measurements made in a few minutes, the question as to the presence of microvariability in RQQSOs has been clearly settled in the affirmative (Gopal-Krishna et al.\\ 2000, 2003; Stalin et al.\\ 2004a,b; Gupta \\& Joshi 2005; Carini et al.\\ 2007; Gupta \\& Yuan 2009; Ram{\\'i}rez et al.\\ 2009). However these papers find that the duty cycle (DC), or fraction of nights such behavior is detected, is low, roughly 10--20\\%, for RQQSOs observed for reasonably long periods (4 hours or more) during a night. On the other hand, blazars show much more rapid variability, with DCs for such monitoring periods around 50--60\\% and when observed for $>6$ hours the blazar DCs rise to 60--85\\% (e.g., Carini 1990; Sagar et al.\\ 2004; Gupta \\& Joshi 2005). The difference in microvariabilty amplitudes and DCs between RQQSOs and blazars can be easily understood if all the fluctuations arise from a jet very close to the central engine, but that jet only escapes that region and emits significant radio power for RL objects (e.g., Gopal-Krishna et al.\\ 2003; Gopal-Krishna, Mangalam \\& Wiita 2008). And most RLQSOs that are not blazars would only have modest (if any) Doppler boosting, so that they also have DCs comparable to those of RQQSOs, as found in the samples of Stalin et al.\\ (2005) and Ram{\\'i}rez et al.\\ (2009), or at a level between those of RQQSOs and blazars (35--40\\%, in the sample of Gupta \\& Joshi 2005), is not surprising. Only recently has a decent sized sample of radio intermediate quasars (RIQs), defined as those relatively rare objects with radio to optical flux ratios in the regime between those that are truly radio quiet and those normally defined as RLQSOs, been targeted for microvariability monitoring (Goyal et al.\\ 2009); their key result is that the RIQs also probably have a DC $\\leq$ 20\\%. Therefore they are not likely to be beamed versions of RQQSOs as has been frequently suggested (Goyal et al.\\ 2009). We conclude that there {\\bf may be a} weak negative correlation between \\hbeta EW and the Eddington ratio, $\\ell$, but there is a {\\bf significant} one between the \\mgii EW and $\\ell$ {\\bf (Fig.~\\ref{lab:lratio_ew_fwhm}; Section~5)}. This latter point has been made independently and more firmly by Dong et al.\\ (2009b) using a larger sample drawn from SDSS. We can also see from Fig.\\ 7 that there is a decline in FWHM with $\\ell$; this is unsurprising since the BH masses are proportional to $L_{edd}$. We see from Fig.\\ 8 that there is a tendency for sources with detectable optical microvariability to have somewhat lower luminosities than those with no such detections. This is interesting, and should be confirmed by examining larger samples. But it is not surprising, as regardless of whether the microvariability arises in accretion discs or jets, more massive BHs have naturally longer characteristic timescales that will presumably make any variations harder to detect over the course of a single night. We also find that the BH masses estimated from the FWHMs of both the \\hbeta and \\mgii lines are reasonable, in that growth to their estimated masses from even small seed BHs is easily possible within the age of the Universe at their observed redshift if the mean $\\ell$ values are close to unity (Tables 4 and 5). This remains true for the great majority of RQQSOs if the value of $\\ell$ we compute from the current continuum flux was constant until the time we observe them; however, this assumption does not work for 4 out of the 19 QSOs with \\hbeta lines. It is also problematical for 1 out of the 26 QSOs with \\mgii lines (i.e., CSO 174, which is also among the difficult set of 4 using \\hbeta profiles). We find no difference between the EWs (of both the \\hbeta or \\mgii lines) and the presence or absence of radio emission in those QSOs (Figs.\\ 4 and 7). As discussed in the introduction, if much of the optical emission in RLQSOs comes from a jet, then we would expect the EWs of the RLQSO sources to be significantly lower than those of the RQQSOs, and if anything, they are slightly higher in our sample. This result does not support the hypothesis (e.g., Czerny et al.\\ 2008) that RQQSOs possess jets that are producing rapid variations. Instead it may indicate that variations involving the accretion disc (e.g., Wiita 2006) play an important role here. Improvements to our results could be obtained through extensive searches for INOV in a larger sample of RQQSOs. It should be useful to divide the sample of sources based on their EWs as available from SDSS, or otherwise uniformly obtained, spectra. Such samples could be divided into three classes based on EW of \\hbeta, e.g., EW$<$40\\AA, 40\\AA $\\le$ EW $<$ 80\\AA\\, and EW $\\ge$ 80\\AA. Such samples should be made as homogeneous as possible on the basis of apparent magnitudes, $z$ and M$_{V}$. With such larger and homogeneous samples, the absence of a correlation between EW and DC of INOV, as found here in our modest sample, could be confirmed or shown to be unlikely." }, "0910/0910.0965_arXiv.txt": { "abstract": "We present the results of X-ray spectral analysis of 22 active galactic nuclei (AGNs) with a small scattering fraction selected from the Second {\\it XMM-Newton} Serendipitous Source Catalogue using hardness ratios. They are candidates of buried AGNs, since a scattering fraction, which is a fraction of scattered emission by the circumnuclear photoionized gas with respect to direct emission, can be used to estimate the size of the opening part of an obscuring torus. Their X-ray spectra are modeled by a combination of a power law with a photon index of 1.5$-$2 absorbed by a column density of $\\sim$ 10$^{23-24}$ cm$^{-2}$, an unabsorbed power law, narrow Gaussian lines, and some additional soft components. We find that scattering fractions of 20 among 22 objects are less than a typical value ($\\sim$ 3\\%) for Seyfert2s observed so far. In particular, those of eight objects are smaller than 0.5\\%, which are in the range for buried AGNs found in recent hard X-ray surveys. Moreover, [\\ion{O}{3}] $\\lambda$5007 luminosities at given X-ray luminosities for some objects are smaller than those for Seyfert2s previously known. This fact could be interpreted as a smaller size of optical narrow emission line regions produced in the opening direction of the obscuring torus. These results indicate that they are strong candidates for the AGN buried in a very geometrically thick torus. ", "introduction": "It is widely accepted that the cosmic X-ray background (CXB) is produced by the integrated emission of faint extragalactic pointlike sources (Brandt \\& Hasinger 2005). {\\it XMM-Newton} and {\\it Chandra} resolved 80\\%$-$100\\% of the CXB at $<$ 2 keV, while the resolved fraction of the CXB at hard X-rays (8$-$12 keV) decreased to only $\\approx$ 50\\% (Worsley et al. 2005). Various observations indicate that a large fraction of active galactic nuclei (AGNs) is hidden by a large amount of cold material (e.g., Awaki et al. 1991; Comastri 2004). According to population synthesis models of the CXB (Comastri et al. 1995; Ueda et al. 2003; Gilli et al. 2007), the peak intensity of the CXB spectrum at 30 keV can be explained by considering contribution of hidden AGNs. Such a population is yet to be understood, because the direct emission from the nucleus is absorbed by surrounding cold gas and is hard to be fully explored with X-ray observations below 10 keV. Hard X-ray surveys performed with {\\it Swift}/BAT (15$-$200 keV; Markwardt et al. 2005; Tueller et al. 2008) and {\\it INTEGRAL} (10$-$100 keV; Bassani et al. 2006; Beckmann et al. 2006; Sazonov et al. 2007) are suitable for unveiling such a type of AGNs with much less selection biases than surveys at lower energies. In fact, AGNs buried in a large amount of matter have been discovered by {\\it Suzaku} follow up observations of a sample selected by the {\\it Swift/}BAT survey (Ueda et al. 2007). In a unified model of an AGN (e.g., Antonucci 1993), torus-like gas is surrounding a supermassive black hole (SMBH), and photoionized gas is created in the opening part of the torus by radiation from the nucleus. If an AGN is observed from the torus side, absorbed direct emission and scattered light by the photoionized gas will be observed in an X-ray spectrum. The fraction of scattered light to direct emission (scattering fraction) can be used to estimate the opening angle of the torus. The fractions for AGNs found by Ueda et al (2007) are extremely small ($<$ 0.5\\%), whereas a typical value is $\\sim$ 3\\% (Turner et al. 1997; Bianchi and Guainazzi 2007). Furthermore, Winter et al (2008) found a similar type of AGNs by {\\it XMM-Newton} observations of {\\it Swift/}BAT selected AGNs. They would be buried in a geometrically thick torus with a very small opening angle assuming that the scattering fraction reflects the solid angle of the opening part of the torus. In an early stage of the evolution of galaxies and their central black holes, a large amount of gas responsible for active star formation may be closely related to obscuration of the nucleus. Therefore, AGNs almost fully covered by an absorber are an important class of objects in studying evolution of AGNs and their hosts. Moreover, they might be significant contributors to the CXB at hard X-rays. Testing a selection technique to find such AGNs and understanding the properties of the population are of significant interest for exploring these issues. We search for buried AGNs with a scattering fraction of 0.5\\% or less using the Second {\\it XMM-Newton} Serendipitous Source Catalogue (2XMM) and the archival data of {\\it XMM-Newton}. We selected candidate sources from the catalogue using hardness ratios (HRs) and scattering fractions calculated by analyzing spectra obtained with {\\it XMM-Newton}. The selection method of candidate sources is described in Section 2. Our results of spectral analysis are presented in Section 3 and their characteristics are discussed in Section 4. Section 5 summarizes our conclusions. We adopt ({\\it H}$_{\\rm 0}$, $\\Omega_{\\rm m}$, $\\Omega_{\\rm \\lambda}$) $=$ (70 km s$^{-1}$Mpc$^{-1}$, 0.3, 0.7) throughout this paper. ", "conclusions": "We searched for AGNs, whose scattered emission is very weak, from the 2XMM Catalogue. In our selection procedure, we calculated HRs expected for an object with a small scattering fraction using a model consisting of absorbed and unabsorbed power laws and 22 sources were selected as candidates from the 2XMM Catalogue. Spectral analysis was conducted using the data observed with {\\it XMM-Newton} for these 22 sources. Their X-ray spectra are represented by a combination of an absorbed power law with a column density of $\\sim$ 10$^{23-24}$ cm$^{-2}$, an unabsorbed power law, a narrow Gaussian line for the Fe K emission, and some additional components. The photon indices of the power-law components for 14 objects are in a typical range of Seyfert 2s ($\\sim$ 1.5$-$2). The photon indices for the others, which have large uncertainties in $\\Gamma$, were fixed at 1.9 in our analysis. We found that the scattering fractions for our sample (except for NGC 4507 and IC 4995) were small compared to a typical value ($\\sim$ 3\\%) of Seyfert 2s observed in the past. In addition, those of eight sources are less than 0.5\\%. If an opening angle of a torus is responsible for the scattering fraction, objects in our sample would be hidden by a geometrically thick torus with a small opening angle. The ratios of {\\it L}$_{\\rm 0.5-2}^{\\rm int}$/{\\it L}$_{\\rm FIR}$ for about a half of our sample are in the range observed for starburst galaxies. This result indicates that thermal emission from a collisionally ionized plasma produced by starburst may contribute to the soft X-ray emission, and true values of scattering fraction for our sample would be smaller than the value calculated in this paper. The distribution of {\\it L}$_{\\rm 2-10}^{\\rm int}$/{\\it L}$_{\\rm [O\\ III]}^{\\rm int}$ for our sample is shifted to higher values than that for other Seyfert 2s studied so far. {\\it L}$_{\\rm [O\\ III]}^{\\rm int}$ depends on the size of the NLR that is considered to be existed in the opening part of the torus. Thus, this result also indicates that the opening angle of the obscuring matter (or scattering optical depth) for our sample is smaller than those for other Seyfert 2, and there is a bias against such a type of AGN buried in a very geometrically thick torus in surveys that rely on optical emission." }, "0910/0910.5547_arXiv.txt": { "abstract": "We describe an implementation of compressible inviscid fluid solvers with block-structured adaptive mesh refinement on Graphics Processing Units using NVIDIA's CUDA. We show that a class of high resolution shock capturing schemes can be mapped naturally on this architecture. Using the method of lines approach with the second order total variation diminishing Runge-Kutta time integration scheme, piecewise linear reconstruction, and a Harten-Lax-van Leer Riemann solver, we achieve an overall speedup of approximately 10 times faster execution on one graphics card as compared to a single core on the host computer. We attain this speedup in uniform grid runs as well as in problems with deep AMR hierarchies. Our framework can readily be applied to more general systems of conservation laws and extended to higher order shock capturing schemes. This is shown directly by an implementation of a magneto-hydrodynamic solver and comparing its performance to the pure hydrodynamic case. Finally, we also combined our CUDA parallel scheme with MPI to make the code run on GPU clusters. Close to ideal speedup is observed on up to four GPUs. ", "introduction": "\\label{sec:introduction} Graphics Processing Units (GPUs) are specialized for math-intensive highly parallel computation, thus more transistors are devoted to data processing rather than data caching and flow control like in CPU. So the potential tremendous performance of general non-graphics computations on GPUs has recently motivated a lot of research activities on general-purpose GPU (GPGPU) computing (see. e.g. Owens et al. 2007 for a review). NVIDIA introduced the Tesla unified graphics and computing architecture in November 2006. The Tesla architecture is built around a scalable array of multithreaded streaming multiprocessors (SM). A SM consists of eight streaming processor (SP) cores. The Tesla SM uses a new processor architecture called single-instruction, multiple-thread (SIMT). The SIMT unit creates, manages and executes up to $768$ concurrent threads in hardware with zero scheduling overhead. The SM also implements barrier synchronization intrinsic with a single instruction. The fast barrier synchronization, together with lightweight thread creation and zero-overhead thread scheduling support very fine-grained parallelism allowing thousands and even millions of threads to be invoked in kernel calls to achieve highly scalable parallel programming. High performance supercomputing has been important in modern astrophysical research since it became available. Simulations allow astronomers to perform ``experiments\" on astronomical objects, collide stars, galaxies, or model the entire visible Universe; all situations are clearly impossible to recreate in a terrestrial laboratory. Studying the formation of stars, black holes and galaxies in the Universe is particularly challenging computationally. Their formation involves the nonlinear interplay of a range of physical processes including gravity, turbulence, magnetic field, shocks, radiation, chemistry, etc. Those questions motivated the astrophysical community to develop robust and efficient fluid codes with all the relevant physics. Studies involving astrophysical fluid dynamics in general are benefitting tremendously from using spatial and temporal adaptive mesh refinement (AMR). This is especially so in the studies of structure formation. For example, the radius of a star is 8 orders of magnitude smaller than the size of a molecular cloud. A uniform grid code is hopeless. On the other hand, the AMR technique has been demonstrated to work well in resolving the large dynamical range involved in those problems (e.g. Abel et al. 2002; Wang \\& Abel 2009). The mapping of computational fluid algorithms to GPU however is still at an early stage of development. Harris et al. (2003) performed cloud simulations using Stam's method (Stam 1999). This method is also used by Liu et al. (2004) for 3D flow calculations. Using finite difference methods, Brandvik \\& Pulla (2008) solved uniform grid 3D Euler equations, Elsen, LeGresley \\& Darve (2008) solved 3D Euler equations on a multi-block meshes and Zink (2008) solved Einstein's equation with uniform grid. As far as we are aware of, this work is the first on mapping an adaptive mesh finite volume solver to GPU. ", "conclusions": "In this work we described how to map HRSC schemes for hyperbolic conservation laws to GPU using NVIDIA's CUDA. We demonstrated that our framework as applied to the equations of inviscid compressible hydrodynamics and MHD can lead to a significant speedup. Specifically, on a Quadro FX 5600 card, saw approximately a factor of ten speedup compared to a single 3 GHz CPU core. An important purpose of our GPU parallel scheme design is achieving good scalability. In typical fluid simulations with as many as millions of cells per grid, our parallelization scheme will launch millions of threads at the same time on GPU to perform the computation. This exceeds the Quadro FX 5600's ability of running $12288$ threads concurrently by more than two orders of magnitude. Thus we expect that the speedup factor of our parallelization scheme will increase linearly with the number of SMs on the GPU. This makes our scheme highly scalable to future generation of graphics hardwares. One important topic we would like to concentrate on in the near future is sparse matrix solvers for Poisson equation. If one can also speed up Poisson solvers by a similar factor on GPU, then coupled with the fluid solvers we implemented in this work, a whole range of astrophysical simulations will be open to processing on the GPU as we can then model astrophysical fluids with self-gravity, viscosity and other non-ideal effects. An implementation of Poisson solver will also make it possible to use projection method for divergence cleaning and implicit fluid solvers." }, "0910/0910.0224_arXiv.txt": { "abstract": "During the last three decades, many papers have reported the existence of a luminosity--metallicity or mass--metallicity ($M$--$Z$) relation for all kinds of galaxies: The more massive galaxies are also the ones with more metal-rich interstellar medium. We have obtained the mass-metallicity relation at different lookback times for the same set of galaxies from the Sloan Digital Sky Survey (SDSS), using the stellar metallicities estimated with our spectral synthesis code {\\sc starlight}. Using stellar metallicities has several advantages: We are free of the biases that affect the calibration of nebular metallicities; we can include in our study objects for which the nebular metallicity cannot be measured, such as AGN hosts and passive galaxies; we can probe metallicities at different epochs of a galaxy evolution. We have found that the $M$--$Z$ relation steepens and spans a wider range in both mass and metallicity at higher redshifts for SDSS galaxies. We also have modeled the time evolution of stellar metallicity with a closed-box chemical evolution model, for galaxies of different types and masses. Our results suggest that the $M$--$Z$ relation for galaxies with present-day stellar masses down to $10^{10} M_\\odot$ is mainly driven by the star formation history and not by inflows or outflows. ", "introduction": " ", "conclusions": "" }, "0910/0910.1222_arXiv.txt": { "abstract": "Although primarily aimed at the galactic archeology and evolution, automated all-sky spectroscopic surveys (RAVE, SDSS) are also a valuable source for the binary star research community. Identification of double-lined spectra is easy and it is not limited by the rare occurrences of eclipses. When the spectrum is properly classified, its atmospheric parameters can be calculated by comparing the spectrum with the best fit atmosphere model. We present the analysis of the binary stars from the sample of roughly 250.000 RAVE survey spectra. The classification and binary discovery method is based on the correlation function analysis. The comparison of these spectra with the model shows that it is possible to estimate the essential atmospheric parameters relatively well. Large number of such estimates and the fact that RAVE consists of a magnitude selected sample without any color cuts makes it suitable for a binary star population study. ", "introduction": "The standard way of solving binary star systems is through the use of sophisticated models, the Wilson-Devinney code for example. This approach is very appealing because it is possible to precisely recover a number of system's parameters. The disadvantage is in the fact that for a reliable solution one needs a large set of photometric as well as spectroscopic observations. The latter only serve as a mean for measuring radial velocities, usually discarding the abundance of other information carried by spectra. It has been proven many times on spectra of single stars \\citep[e.g.~][to name just a few]{bailer-jones1997,katz1998,decin2004,ocvirk2006,koleva2009} that it is possible to estimate some of the astrophysical parameters (e.g.~temperature, gravity, and metallicity) by fitting the observed spectrum with the stellar atmosphere model. The same can be done in case of binary stars. This approach might be particularly useful in the pipelines of spectroscopic surveys, where parameters of large amount of observed objects have to be extracted automatically. In the following section we will briefly review the RAVE survey itself and present the results of estimation of the most significant parameters from the spectra of binary stars in the RAVE survey. Before that we will also deal with another problem that has to be addressed when working with large datasets of unknown objects, the discovery of binary stars. Because it is not known in advance if some spectrum belongs to a binary star or some other type of peculiar object, an automatic method has to be used to properly classify all spectra. ", "conclusions": "The described classification and parameter estimation methods gave satisfactory results to some extent, but there is still room for improvement. Other methods like the mentioned nested sampling seem very promising for such tasks. Also, the planned use of stellar evolution models will additionally constrain the errors of calculated parameters, yielding better results. The forthcoming missions like Gaia and the Hermes survey, will provide even more data with higher resolution. The need for truly automatic processing pipelines will be even grater, so it is important to gain as much experience as possible until then." }, "0910/0910.4541_arXiv.txt": { "abstract": "\\noindent Using a Thomas-Fermi model, we calculate the structure of the electrosphere of the quark antimatter nuggets postulated to comprise much of the dark matter. This provides a single self-consistent density profile from ultrarelativistic densities to the nonrelativistic Boltzmann regime that use to present microscopically justified calculations of several properties of the nuggets, including their net charge, and the ratio of MeV to 511~keV emissions from electron annihilation. We find that the calculated parameters agree with previous phenomenological estimates based on the observational supposition that the nuggets are a source of several unexplained diffuse emissions from the Galaxy. As no phenomenological parameters are required to describe these observations, the calculation provides another nontrivial verification of the dark-matter proposal. The structure of the electrosphere is quite general and will also be valid at the surface of strange-quark stars, should they exist. ", "introduction": "\\label{sec:introduction}\\noindent In this paper we explore some details of a testable and well-constrained model for dark matter~\\cite{Zhitnitsky:2002qa, Oaknin:2003uv, Zhitnitsky:2006vt, Forbes:2006ba, Forbes:2008uf} in the form of quark matter as antimatter nuggets. In particular, we focus on physics of the ``electrosphere'' surrounding these nuggets: It is from here that observable emissions emanate, allowing for the direct detection of these dark-antimatter nuggets. We first provide a brief review of our proposal in Sec.~\\ref{sec:proposal}, then describe the structure of the electrosphere of the nuggets using a Thomas-Fermi model in Sec.~\\ref{sec:electr-struct}. This allows us to calculate the charge of the nuggets, and to discuss how they maintain charge equilibrium with the environment. We then apply these results to the calculation of emissions from electron annihilation in Sec.~\\ref{sec:diffuse-1-20}, computing some of the phenomenological parameters introduced in~\\cite{Lawson:2007kp} required to explain current observations. The values computed in the present paper are consistent with these phenomenologically motivated values, providing further validation of our model for dark matter. The present results concerning the density profile of the electrosphere may also play an important role in the study of the surface of quark stars, should they exist. ", "conclusions": "\\noindent Solving the relativistic Thomas-Fermi equations, we determined the charge and structure of the positron electrosphere of quark antimatter nuggets that we postulate could comprise the missing dark-matter in our Universe. We found the structure of the electrosphere to be insensitive to the size of the nuggets, as long as they are large enough to be consistent with current terrestrial based detector limits, and hence, can make unambiguous predictions about electron annihilation processes. To test the dark-matter postulate further, we used the structure of this electrosphere to calculate the annihilation spectrum for incident electronic matter. The model predicts two distinct components: a 511 keV emission line from decay through a positronium intermediate and an MeV continuum emission from direct-annihilation processes deep within the electrosphere. By fixing the general normalization to the measured 511 keV line intensity seen from the core of the Galaxy, our model makes a definite prediction about the intensity and spectrum of the MeV continuum spectrum without any additional adjustable model parameters: Our predictions are based on well-established physics. As discussed in~\\cite{Lingenfelter:2009fk}, a difficulty with most other dark-matter explanations for the 511 keV emission is to explain the large $\\sim$100\\% observed positronium fraction -- positrons produced in hot regions of the Galaxy would produce a much smaller fraction. Our model naturally predicts this observed ratio everywhere. \\textit{A priori}, there is no reason to expect that the predicted MeV spectrum should correspond to observations: typically two uncorrelated emissions are separated by many orders of magnitude. We find that the phenomenological parameter $\\chi$~(\\ref{eq:branching}) required to explain the relative normalization of MeV emissions arises naturally from our microscopic calculation. This is highly nontrivial because it requires a delicate balance between the two annihilation processes from the semirelativistic region of densities that is sensitive to the semirelativistic self-consistent structure of the electrosphere outside the range of validity of the analytic ultrarelativistic and nonrelativistic regimes. (See Fig.~\\ref{fig:antinugget_profiles} and~\\ref{fig:survival}). If the predicted emission were several orders of magnitude too large, the observations would have ruled out our proposal. If the predicted emissions were too small, the proposal would not have been ruled out, but would have been much less interesting. Instead, we are left with the intriguing possibility that both the 511 keV spectrum and much of the MeV continuum emission arise from the annihilation of electrons on dark-antimatter nuggets. While not a smoking gun -- at least until the density and velocity distributions of matter and dark-matter are much better understood -- this provides another highly nontrivial test of the proposal that, \\textit{a priori}, could have ruled it out. Both the formal calculations and the resulting structure presented here -- spanning density regimes from ultrarelativistic to nonrelativistic -- are similar to those relevant to electrospheres surrounding strange-quark stars should they exist. Therefore, our results may prove useful for studying quark star physics. In particular, problems such as bremsstrahlung emission from quark stars originally analyzed in~\\cite{Jaikumar:2004rp} (and corrected in~\\cite{Caron:2009zi}) that uses only ultrarelativistic profile functions. The results of this work can be used to generalize the corresponding analysis for the entire range of allowed temperatures and chemical potentials. Another problem which can be analyzed using the results of the present work is the study of the emission of energetic electrons produced from the interior of quark stars. As advocated in~\\cite{Charbonneau:2009ax}, these electrons may be responsible for neutron star kicks, helical and toroidal magnetic fields, and other important properties that are observed in a number of pulsars, but are presently unexplained. Finally, we would like to emphasize that this mechanism demonstrates that dark matter may arise from \\emph{within} the standard model at the \\QCD\\ scale,\\footnote{The axion is another dark-matter candidate arising from the \\QCD\\ scale, but with fewer observational consequences.} and that exotic new physics is not required. Indeed, this is naturally suggested by the ``cosmic coincidence'' of almost equal amounts of dark and visible contributions to the total density $\\Omega_{\\text{tot}} = 1.011(12)$ of our Universe~\\cite{Amsler:2008zz}: \\begin{equation*} \\Omega_{\\text{dark-energy}} : \\Omega_{\\text{dark-matter}} : \\Omega_{\\text{visible}} \\approx 17:5:1. \\end{equation*} The dominant baryon contribution to the visible portion $\\Omega_{\\text{visible}} \\approx \\Omega_{B}$ has an obvious relation to \\QCD\\ through the nucleon mass $m_{\\textsc{n}} \\propto \\Lambda_{\\QCD}$ (the actual quark masses arising from the Higgs mechanism contribute only a small fraction to $m_{\\textsc{n}}$). Thus, a \\QCD\\ origin for the dark components would provide a natural solution to the extraordinary ``fine-tuning'' problem typically required by exotic high-energy physics proposals. Our proposal here solves the matter portion of this coincidence. For a proposal addressing the energy coincidence we refer the reader to \\cite{Urban:2009vy,*Urban:2009yg} and references therein.\\footnote{The idea concerns the anomaly that solves the famous axial $U(1)_{A}$ problem, giving rise to an $\\eta'$ mass that remains finite, even in the chiral limit. Under some plausible and testable assumptions about the topology of our Universe, the anomaly demands that the cosmological vacuum energy depend on the Hubble constant $H$ and \\QCD\\ parameters as $\\rho_{\\textsc{de}} \\sim H m_q\\langle\\bar{q}q\\rangle /m_{\\eta'} \\approx (4 \\times 10^{-3}\\text{ eV})^4$ -- tantalisingly close to the value $\\rho_{\\textsc{de}} = [1.8(1)\\times 10^{-3}\\text{ eV}]^4$ observed today~\\cite{Amsler:2008zz}.}" }, "0910/0910.1478_arXiv.txt": { "abstract": "In this work we study the internal spatial structure of 16 open clusters in the Milky Way spanning a wide range of ages. For this, we use the minimum spanning tree method (the $Q$ parameter, which enables one to classify the star distribution as either radially or fractally clustered), King profile fitting, and the correlation dimension ($D_c$) for those clusters with fractal patterns. On average, clusters with fractal-like structure are younger than those exhibiting radial star density profiles. There is a significant correlation between $Q$ and the cluster age measured in crossing time units. For fractal clusters there is a significant correlation between the fractal dimension and age. These results support the idea that stars in new-born clusters likely follow the fractal patterns of their parent molecular clouds, and eventually evolve toward more centrally concentrated structures. However, there can exist stellar clusters as old as $\\sim 100$ Myr that have not totally destroyed their fractal structure. Finally, we have found the intriguing result that the lowest fractal dimensions obtained for the open clusters seem to be considerably smaller than the average value measured in galactic molecular cloud complexes. ", "introduction": "The hierarchical structure observed in some open clusters is presumably a consequence of its formation in a turbulent medium with an underlying fractal structure (\\cite[Elmegreen \\& Scalo 2004]{Elm04}). Otherwise, open clusters having central star concentrations with radial star density profiles likely reflect the dominant role of gravity, either on the primordial gas structure or as a result of a rapid evolution from a more structured state (\\cite[Lada \\& Lada 2003]{Lad03}). Therefore, the analysis of the distribution of stars may yield information on the formation process and early evolution of open clusters. It is necessary, however, that this kind of analysis is done by measuring the cluster structure in an objective, quantitative, as well as systematic way. Here we study the internal spatial structure in a sample of 16 open clusters spanning a wide range of ages. ", "conclusions": "Our results support the idea that stars in new-born cluster likely follow the fractal patterns of their parent molecular clouds, and that eventually evolve toward more centrally concentrated structures (see \\cite[Schemja \\& Klessen 2006]{Sch06}; \\cite[Schmeja et al. 2008]{Sch08}, \\cite[2009]{Sch09}; \\cite[S\\'anchez et al. 2007a]{San07a}, \\cite[2009]{San09}). However, this seems to be only an overall trend. The very young cluster $\\sigma$ Orionis (age $\\sim 3$ Myr) exhibits a radial density gradient with $Q \\simeq 0.88$ (\\cite[Caballero 2008]{Cab08}). On the other hand, Table~\\ref{data} shows open clusters as old as $\\sim 100$ Myr that have not totally destroyed their clumpy structure (for example, both NGC~1513 and NGC~1647 have $Q \\sim 0.7$). \\cite[Goodwin \\& Whitworth (2004)]{Goo04} simulated the dynamical evolution of young clusters and showed that the survival of the initial substructure depends strongly on the initial velocity dispersion. Fractal clusters with a low velocity dispersion tend to erase their substructure rather quickly. However, if the velocity dispersion is high, such that the cluster remains supported against its own gravity or even expands, then significant levels of substructure can survive for several crossing times. Thus, our results give some observational support to \\cite[Goodwin \\& Whitworth's (2004)]{Goo04} simulations. From Fig.~\\ref{correlaDc}, we can see that clusters with the smallest correlation dimensions ($D_c = 1.74$) would have three-dimensional fractal dimensions around $D_f \\sim 2.0$ (estimated from previous papers, see e.g. Fig.~1 in \\cite[S\\'anchez \\& Alfaro 2008]{San08}). This is a very interesting result because this value is considerably smaller than the average value estimated for galactic molecular clouds in recent studies, which is $D_f \\simeq 2.6-2.7$ (\\cite[S\\'anchez et al. 2005]{San05}, \\cite[S\\'anchez et al. 2007b]{San07b}). Young, new-born stars probably will reflect the conditions of the interstellar medium from which they were formed. Therefore, a group of stars born from the same cloud, i.e. born at almost the same place and time, should have a fractal dimension similar to that of the parent cloud. If the fractal dimension of the interstellar medium has a nearly universal value around 2.6-2.7, then how can some clusters exhibit such small fractal dimension values? Perhaps some clusters may develop some kind of substructure starting from an initially more homogeneous state. This possibility has been confirmed in numerical simulations (\\cite[Goodwin \\& Whitworth 2004]{Goo04}), although some coherence in the initial velocity dispersion is required. Another explanation is that this difference is a consequence of a more clustered distribution of the densest gas from which stars form at the smallest spatial scales in the molecular cloud complexes, according to a multifractal scenario (\\cite[Chappell \\& Scalo 2001]{Cha01}). Maybe the star formation process itself modifies in some (unknown) way the underlying geometry generating distributions of stars that can be very different from the distribution of gas in the parental clouds. Finally, one possibility is that the fractal dimension of the interstellar medium in the Galaxy does not have a universal value and therefore some regions form stars distributed following more clustered patterns. There is no a priori reason for assuming that $D_f$ has nearly the same value everywhere in the Galaxy independently of either the dominant physical processes or environmental variables. Recent simulations of supersonic isothermal turbulence done by \\cite[Federrath et al. (2009)]{Fed09} showed that compressive forcing yields fractal dimension values for the interstellar medium significantly smaller ($D_f \\sim 2.3$) compared to solenoidal forcing ($D_f \\sim 2.6$). Thus, $D_f$ could be very different from region to region in the Galaxy depending on the main physical processes driving the turbulence. At least at galactic scales, it has been shown that there are significant differences in the fractal dimension of the distribution of star forming sites among the galaxies, contrary to the universal picture previously claimed in the literature (\\cite[see S\\'anchez \\& Alfaro 2008]{San08}). So that the possibility of a non-universal fractal dimension for the interstellar medium in the Galaxy cannot, in principle, be ruled out. \\begin{figure}[t] \\begin{minipage}[t]{0.48\\linewidth} \\includegraphics[scale=0.52]{correlaQ.eps} \\caption{Structure parameter $Q$ as a function of the logarithm of age divided by the tidal radius, which is nearly proportional to age in crossing times units. The dashed line at $Q=0.8$ roughly separates radial from fractal clustering. The best linear fit (equation at the top) is represented by a solid line.} \\label{correlaQ} \\end{minipage} \\hfill \\begin{minipage}[t]{0.48\\linewidth} \\includegraphics[scale=0.52]{correlaDc.eps} \\caption{Calculated correlation dimension as a function of age (in crossing time units). The best linear fit (equation at the top) is represented by a solid line. As reference, horizontal dashed lines indicate the values corresponding to three-dimensional distributions with fractal dimensions of $D_f=2.0$ and $2.5$.} \\label{correlaDc} \\end{minipage} \\end{figure}" }, "0910/0910.4294_arXiv.txt": { "abstract": "{Although the {{\\it internal linear combination}} method (ILC) is a technique widely used for the separation of the Cosmic Microwave Background signal from the Galactic foregrounds, its characteristics are not yet well defined. This can lead to misleading conclusions about the actual potentialities and limits of such approach in real applications. Here we discuss briefly some facts about ILC that to our knowledge are not fully worked out in literature and yet have deep effects in the interpretation of the results.} ", "introduction": "A widely used approach for the separation of the {\\it cosmic microwave background} (CMB) from the diffuse Galactic background is the {\\it internal linear combination} method (ILC). For instance, this method was adopted in the reduction of the data from the Wilkinson Microwave Anisotropy Probe (WMAP) satellite for CMB observations \\citep{ben03}. Its success is due to the fact that, among the separation techniques, ILC calls for the smallest number of {\\it a priori} assumptions. If the data are in the form of $N_o$ maps, taken at different frequencies and containing $N_p$ pixels each, the model on which ILC is based is \\begin{equation} \\label{eq:image} \\Si = \\Scmb + \\Sgal + \\Nmthi(p). \\end{equation} Here, $\\Si$ provides the value of the $p$th pixel for a map obtained at channel ``$~i~$'' \\footnote{In the present work, $p$ indexes pixels in the classic spatial domain. However, the same formalism applies if other domains are considered, for example, the Fourier one.}, whereas $\\Scmb$, $\\Sgal$ and $\\Nmthi(p)$ are the contributions due to the CMB, the diffuse Galactic foreground and the experimental noise, respectively. Although not necessary, often it is assumed that all of these contributions are representable by means of stationary random fields. Moreover, without loss of generality, for ease of notation the random fields are supposed as the realization of zero-mean spatial processes. The basic idea behind model~(\\ref{eq:image}) is that, contrary to the components that form the Galactic background, CMB is independent of the observing channel. ILC exploits this fact averaging $N_o$ images $\\{ \\Si \\}_{i=1}^{N_o}$ and giving a specific weight $w_i$ to each of them so as to minimize the impact of the foreground and noise \\citep{ben03}. This means to look for a solution of type \\begin{equation} \\label{eq:sum} \\Scmbh = \\sum_{i=1}^{N_o} w_i \\Si. \\end{equation} If the constraint $\\sum_{i=1}^{N_o} w_i = 1$ is imposed, Eq.~(\\ref{eq:sum}) becomes \\begin{equation} \\label{eq:wls} \\Scmbh = \\Scmb + \\sum_{i=1}^{N_o} w_i [ \\Sgal + \\Nmthi(p) ]. \\end{equation} Now, from this equation it is clear that, for a given pixel ``$p$'', the only variable terms are in the sum. Hence, under the assumption of independence of $\\Scmb$ from $\\Sgal$ and $\\Nmthi(p)$, the weights $\\{ w_i \\}$ have to minimize the variance of $\\Scmbh$, i.e. \\begin{align} & \\{ w_i \\} = \\underset{\\{ w_i \\} }{\\arg\\min} \\nonumber \\\\ & {\\rm VAR} \\left[\\Scmb \\right] + {\\rm VAR}\\left[\\sum_{i=1}^{N_o} w_i (\\Sgal + \\Nmthi(p)) \\right], \\end{align} where ${\\rm VAR}[s(p)]$ is the {\\it expected variance} of $s(p)$. If $\\Sib$ denotes a {\\bf row vector} such as $\\Sib = [S^{(i)}(1), S^{(i)}(2), \\ldots, S^{(i)}(N_p)]$ and the $N_o \\times N_p$ matrix $\\Sb$ is defined as \\begin{equation} \\Sb = \\left( \\begin{array}{c} \\Sb^{(1)} \\\\ \\Sb^{(2)} \\\\ \\vdots \\\\ \\Sb^{(N_o)} \\end{array} \\right), \\end{equation} then Eq.~(\\ref{eq:image}) becomes \\begin{equation} \\label{eq:basicmm} \\Sb = \\Scmbb + \\Sgalb + \\Nmthb. \\end{equation} In this case, the weights are given by \\citep{eri04} \\begin{equation} \\label{eq:wr} \\wb = \\frac{\\Cb_{\\Sb}^{-1} \\oneb}{\\oneb^T \\Cb_{\\Sb}^{-1} \\oneb}, \\end{equation} where $\\Cb_{\\Sb}$ is the $N_o \\times N_o$ cross-covariance matrix of the random processes that generate $\\Sb$, i.e. \\begin{equation} \\Cb_{\\Sb} = {\\rm E}[\\Sb \\Sb^T], \\end{equation} and $\\oneb = (1, 1, \\ldots, 1)^T$ is a column vector of all ones. Here, ${\\rm E}[.]$ denotes the {\\it expectation operator}. Hence, the ILC estimator takes the form \\begin{align} \\Scmbhb & = \\wb^T \\Sb, \\label{eq:wlss} \\\\ & = \\alpha \\oneb^T \\Cb_{\\Sb}^{-1} \\Sb, \\label{eq:basic} \\end{align} with $\\oneb^T \\wb = 1$ and the scalar quantity $\\alpha$ given by \\begin{equation} \\label{eq:alpha} \\alpha = [ \\oneb^T \\Cb_{\\Sb}^{-1} \\oneb]^{-1}. \\end{equation} In practical applications, matrix $\\Cb_{\\Sb}$ is unknown and has to be estimated from the data. Typically, this is done by means of the estimator \\begin{equation} \\label{eq:C} \\widehat \\Cb_{\\Sb} = \\frac{1}{N_p} \\Sb \\Sb^T. \\end{equation} In this case, the ILC estimator is given by Eqs.(\\ref{eq:wlss})-(\\ref{eq:alpha}) with $\\Cb_{\\Sb}$ and $\\wb$ replaced, respectively, by $\\widehat \\Cb_{\\Sb}$ and \\begin{equation} \\label{eq:w} \\widehat \\wb = \\frac{\\widehat \\Cb_{\\Sb}^{-1} \\oneb}{\\oneb^T \\widehat \\Cb_{\\Sb}^{-1} \\oneb}. \\end{equation} ", "conclusions": "From the analysis presented above, it is evident that the results provided by ILC have to be interpreted with extreme caution. This techniques suffers many drawbacks that, if not properly taken into account, can lead to misleading if not wrong conclusions. In particular, ILC should be used only in situations where matrix $\\widehat\\Cb_{\\Sb}$ is (close to be) singular, i.e. only when one can be certain that model~(\\ref{eq:model})-(\\ref{eq:Amatrix}) holds with $N_o > N_c+1$. In the contrary case, the separation operated through ILC has to be expected inaccurate." }, "0910/0910.3361_arXiv.txt": { "abstract": "We studied how the intergalactic magnetic field (IGMF) affects the propagation of super-GZK protons that originate from extragalactic sources within the local GZK sphere. Toward this end, we set up hypothetical sources of ultra-high-energy cosmic-rays (UHECRs), virtual observers, and the magnetized cosmic web in a model universe constructed from cosmological structure formation simulations. We then arranged a set of reference objects mimicking active galactic nuclei (AGNs) in the local universe, with which correlations of simulated UHECR events are analyzed. With our model IGMF, the deflection angle between the arrival direction of super-GZK protons and the sky position of their actual sources is quite large with the mean value of $\\langle \\theta \\rangle \\sim 15^{\\circ}$ and the median value of $\\tilde \\theta \\sim 7 - 10^{\\circ}$. On the other hand, the separation angle between the arrival direction and the sky position of nearest reference objects is substantially smaller with $\\langle S \\rangle \\sim 3.5 - 4^{\\circ}$, which is similar to the mean angular distance in the sky to nearest neighbors among the reference objects. This is a direct consequence of our model that the sources, observers, reference objects, and the IGMF all trace the matter distribution of the universe. The result implies that extragalactic objects lying closest to the arrival direction of UHECRs are not necessary their actual sources. With our model for the distribution of reference objects, the fraction of super-GZK proton events, whose closest AGNs are true sources, is less than 1/3. We discussed implications of our findings for correlation studies of real UHECR events. ", "introduction": "The nature and origin of ultra-high-energy cosmic rays (UHECRs), especially above the so-called Greisen-Zatsepin-Kuz'min (GZK) energy of $E_{\\rm GZK}\\approx$ 50 EeV (1 EeV $=10^{18}$ eV), has been one of most perplexing puzzles in astrophysics over five decades and still remains to be understood \\citep[see][for a review]{nw00}. The highest energy CR detected so far is the Fly's Eye event with an estimated energy of $\\sim 300$ EeV \\citep{flyseye94}. At these high energies protons and nuclei cannot be confined and accelerated effectively within our Galaxy, so the sources of UHECRs are likely to be extragalactic. At energies higher than $E_{\\rm GZK}$, it is expected that protons lose energy and nuclei are photo-disintegrated via the interactions with the cosmic microwave background radiation (CMB) along their trajectories in the intergalactic space \\citep{grei66,zk66,psb76}. The former is known as the GZK effect. So a significant suppression in the energy spectrum above $E_{\\rm GZK}$ could be regarded as an observational evidence for the extragalactic origin of UHECRs \\citep[see, e.g.,][]{bgg06}. However, the accurate measurement of the UHECR spectrum is very difficult, partly because of extremely low flux of UHECRs. But a more serious hurdle is the uncertainties in the energy calibration inherent in detecting and modeling extensive airshower events \\citep[e.g.,][]{nw00, watson06}. Nevertheless, both the Yakutsk Extensive Air Shower Array (Yakutsk) and the High Resolution Fly's Eyes (HiRes) reported observations of the GZK suppression \\citep{yakutsk04,hires08a}, while the Akeno Giant Air Shower Array (AGASA) claimed a conflicting finding of no suppression \\citep{agasa04}. More recent data from the Pierre Auger Observatory (Auger) support the existence of the GZK suppression \\citep{auger08b,auger09c}. Below $E_{\\rm GZK}$, however, the four experiments reported the fluxes that are different from each other by up to a factor of several, implying the possible existence of systematic errors in their energy calibrations \\citep{bere09}. The overall sky distribution of the arrival directions of UHECRs below $E_{GZK}$ seems to support the isotropy hypothesis \\citep[see, e.g.,][]{nw00}. This is consistent with the expectation of uniform distribution of extragalactic sources; the interaction length (i.e. horizon distance) of protons below $E_{GZK}$ is a few Gpc and the universe can be considered homogeneous and isotropic on such a large scale. The horizon distance for super-GZK events, however, decreases sharply with energy and $R_{\\rm GZK} \\sim 100 $ Mpc for $E=100$EeV \\citep{bg88}. The matter distribution inside the local GZK horizon ($R_{\\rm GZK}$) is inhomogeneous. Since powerful astronomical objects are likely to form at deep gravitational potential wells, we expect the distribution of the UHECR sources would be inhomogeneous as well. Hence, if super-GZK proton events point their sources, their arrival directions should be anisotropic. The anisotropy of super-GZK events, hence, has been regarded to provide an important clue that unveils the sources of UHECRs. So far, however, the claims derived from analyses of different experiments are often tantalizing and sometimes conflicting. For instance, with an excessive number of pairs and one triplet in the arrival direction of CRs above 40 EeV, the AGASA data support the existence of small scale clustering \\citep{agasa96,agasa99}. On the other hand, the HiRes stereo data are consistent with the hypothesis of null clustering \\citep{hires04,hires09}. The auto-correlation analysis of the Auger data reported a weak excess of pairs for $E > 57$ EeV \\citep{auger08a}. In addition, the Auger Collaboration found a correlation between highest energy events and the large scale structure (LSS) of the universe using nearby active galactic nuclei (AGNs) in the \\citet{vcv06} catalog \\citep{auger07a,auger08a,auger09b} as well as using nearby objects in different catalogs \\citep{auger09a}. A correlation between highest AGASA events with nearby galaxies from SDSS was reported \\citep{tns09}. The HiRes data, however, do not show such correlation of highest energy events with nearby AGNs \\citep{hires08b}, but instead show a correlation with distant BL Lac objects \\citep{hires06}. The interpretation of anisotropy and correlation analyses is, however, complicated owing to the intervening galactic magnetic field (GMF) and intergalactic magnetic field (IGMF); the trajectories of UHECRs are deflected by the magnetic fields as they propagate through the space between sources and us, and hence, their arrival directions are altered. Even with considerable observational and theoretical efforts, however, the nature of the GMF and the IGMF is still poorly constrained. Yet, models for the GMF generally assume a strength of $\\sim$ a few $\\mu$G and a coherence length of $\\sim 1$ kpc for the field in the Galactic halo \\citep[see, e.g.,][]{stan97}, and predict the deflection of UHE protons due to the GMF to be $\\theta \\sim$ a few degrees \\citep[see, e.g.,][]{ts08}. The situation for the IGMF in the LSS has been confusing. Adopting a model for the IGMF with the average strength of $\\langle B \\rangle \\sim 100$ nG in filaments, \\citet{sme03} showed that the deflection of UHECRs due to the IGMF could be very large, e.g., $\\theta > 20^{\\circ}$ for protons above 100 EeV. On the other hand, \\citet{dgst05} adopted a model with $\\langle B \\rangle \\sim 0.1$ nG in filaments and showed that the deflection should be negligible, e.g., $\\theta \\ll 1^{\\circ}$ for protons with 100 EeV. Recently, \\citet{rkcd08} proposed a physically motivated model for the IGMF, in which a part of the gravitational energy released during structure formation is transferred to the magnetic field energy as a result of the turbulent dynamo amplification of weak seed fields in the LSS of the universe. In the model, the IGMF follows largely the matter distribution in the cosmic web, and the strength and coherence length are predicted to be $\\langle B \\rangle \\sim 10$ nG and $\\sim 1$ Mpc for the field in filaments. Such field in filaments is expected to induce the Faraday rotation \\citep{cr09}, which is consistent with observation \\citep{xkhd06}. With this model IGMF, \\citet{dkrc08} (Paper I hereafter) calculated the trajectories of UHE protons ($E>10$ EeV) that were injected at extragalactic sources associated with the LSS in a simulated model universe. We then estimated that only $\\sim 35$ \\% of UHE protons above 60 EeV would arrive at us with $\\theta \\leq 5^{\\circ}$ and the average value of deflection angle would be $\\langle \\theta \\rangle \\sim 15^{\\circ}$. Note that the deflection angle of $\\langle \\theta \\rangle \\sim 15^{\\circ}$ is much larger than the angular window of $3.1^{\\circ}$ used by the Auger collaboration in the study of the correlation between highest energy UHECR events and nearby AGNs \\citep{auger07a,auger08a,auger09b}. In this contribution, as a follow-up work of Paper I, we investigate the effects of the IGMF on the arrival direction of super-GZK protons above 60 EeV coming from sources within 75 Mpc. The limiting parameters for energy and source distance are chosen to match the recent analysis of the Auger collaboration. Without knowing the true sources of UHECRs, the statistics that can be obtained with observational data from experiments are limited; some statistics that are essential to reveal the nature of sources are difficult or even impossible to be constructed. On the other hand, with data from simulations, any statistics can be explored. In that sense, simulations complement experiments. Here, with the IGMF suggested by \\citet{rkcd08}, we argue that the large deflection angle of super-GZK protons due to the IGMF is not inconsistent with the anisotropy and correlation recently reported by the Auger collaboration. However, the large deflection angle implies that the nearest object to a UHECR event in the sky is not necessarily its actual source. In Section 2, we describe our models for the LSS of the universe, IGMF, observers, and sources of UHECRs, reference objects for correlation study, and simulations. In Section 3, we present the results, followed by a summary and discussion in Section 4. ", "conclusions": "In the search for the nature and origin of UHECRs, understanding the propagation of charged particles through the magnetized LSS of the universe is important. At present, the details of the IGMF are still uncertain, mainly due to limited available information from observation. Here, we adopted a realistic model universe that was described by simulations of cosmological structure formation; our simulated universe represents the LSS, which is dominated by the cosmic web of filaments interconnecting clusters and groups. The distribution of the IGMF in the LSS of the universe was obtained with a physically motivated model based on turbulence dynamo \\citep{rkcd08}. To investigate the effects of the IGMF on the arrival direction of UHECRs, we further adopted the following models. Virtual observers of about 1400 were placed at groups of galaxies, which represent statistically the Local Group in the simulated model universe. Then, we set up a set of about 500 AGN-like ``reference objects'' within 75 Mpc from each observer, at clusters of galaxies (deep gravitational potential wells) along the LSS. They represent a class of astronomical objects with which we performed a correlation analysis for simulated UHECR events. We considered three models, in which subsets of the reference objects were selected as AGN-like sources of UHECRs (see Table 1). UHE protons of $E \\ge 60$ EeV with power-low energy spectrum were injected at those sources, and the trajectories of UHE protons in the magnetized cosmic web were followed. At observer locations, the events with $E \\ge 60$ EeV from sources within a sphere of radius 75 Mpc were recorded and analyzed. To characterize the clustering of the reference objects, we calculated the angular distance, $Q$, from a given reference object to its nearest neighbor. The mean value for our model AGNs in the simulated universe is $\\langle Q_{\\rm AGN} \\rangle = 3.68^{\\circ} \\pm 1.66^{\\circ}$, while that for 442 AGNs from the VCV catalog is $\\langle Q_{\\rm VCV} \\rangle = 3.55^{\\circ}$. This demonstrates that the two samples have a similar degree of clustering and are highly structured (e.g. $\\langle Q_{\\rm iso}\\rangle \\approx 11^{\\circ}$ for the isotropic distribution). With our model IGMF, the deflection angle, $\\theta$, between the arrival direction of UHE protons and the sky position of their actual sources, is quite large with the mean $\\langle \\theta \\rangle \\sim 14 - 17.5^{\\circ}$ and the median $\\tilde \\theta \\sim 7 - 10^{\\circ}$, depending on models with different numbers of sources (see Table 1). On the other hand, the separation angle between the arrival direction and the sky position of nearest reference objects is substantially smaller with the mean $\\langle S_{\\rm sim} \\rangle \\sim 3.5 - 4^{\\circ}$ and the median $\\tilde S_{\\rm sim} = 2.8 - 3.5^{\\circ}$. That is, we found that while $\\langle \\theta \\rangle \\sim 4 \\langle S_{\\rm sim} \\rangle$, $\\langle S_{\\rm sim} \\rangle$ is similar to $\\langle Q_{\\rm AGN} \\rangle$. For the Auger events of highest energies in \\citet{auger08a}, with 442 nearby AGNs from the VCV catalog as the reference objects, the mean separation angle is $\\langle S_{\\rm Auger} \\rangle = 3.23^{\\circ}$ for the 26 events, excluding one event with large $S$ ($ \\approx 27^{\\circ}$), while $\\langle S_{\\rm Auger} \\rangle =4.13^{\\circ}$ for all the 27 events. Hence, $\\langle S_{\\rm Auger} \\rangle \\sim \\langle Q_{\\rm VCV} \\rangle$ $\\sim \\langle S_{\\rm sim} \\rangle \\sim \\langle Q_{\\rm AGN} \\rangle $. This implies that the separation angle from the Auger data would be determined primarily by the distribution of reference objects (AGNs), and may not represent the true deflection angle. We further tested whether the distributions of separation angle, $S$, for our simulated events and for the Auger events are statistically comparable to each other. According the Kolmogorov-Smirnov test for the cumulative fraction of events, $F(\\le \\log S)$, versus $\\log S$, the significance level of the null hypothesis that the two distributions are drawn from the identical population is as large as $P\\sim 0.37$ for Model A (see Table 1). Thus, we argued that our simulation data, especially in Model A, are in a fair agreement with the Auger data. This test also showed that the model with more sources (Model A) is preferred over the models with fewer sources (Models B and C). The fact that $\\langle \\theta \\rangle \\gg \\langle S_{\\rm sim} \\rangle$ implies that the AGNs found closest to the direction of UHECRs may not be the true sources of UHECRs. We estimated the probability of finding the true sources of UHECRs, when nearest reference objects are blindly chosen: $f(S)$ is the ratio of the number of true source identifications to the total number of simulated events. This probability is $\\sim 50 - 30$ \\% at $S \\sim 2^{\\circ}$, but decreases to $\\sim 10$ \\% at larger separation angle. On average, in less than 1 out of 3 events, the true sources of UHECRs can be identified in our simulations, when nearest reference objects are chosen. The distribution of $\\theta$ versus $D_{\\theta}$ shows a bimodal pattern in which $\\theta$ is on average larger either for nearby sources (for $D_{\\theta} \\la 15$ Mpc) or for distant sources (for $D_{\\theta} \\ga 30$ Mpc) with the minimum at the intermediate distance of $D_{\\theta,{\\rm min}} \\sim 20 - 30$ Mpc. The distribution of $S$ versus $D_s$ shows a similar, but weaker sign of the bimodal pattern. This behavior is a characteristic signature of the magnetized cosmic web of the universe, where filaments are the most dominant structure. When a large number of super-GZK events are accumulated, we may find the signature of the cosmic web of filaments in the $S$ versus $D_s$ distribution. Finally, we address the limitations of our work. (1) We worked in a simulated universe with specific models for the elements such as the IGMF, observers, sources, and reference objects, but not in the real universe. So we could make only statistical statements. (2) It has been shown previously that adopting different models for the IGMF, very different deflection angles are obtained \\citep[see][]{sme03,dgst05,dkrc08}. We argue that our model for the IGMF is most plausible, since it is a physically motivated model based on turbulence dynamo without involving an arbitrary normalization \\citep{rkcd08}. Nevertheless, our IGMF model should be confirmed further by observation. (3) The sources of UHECRs may not be objects like AGNs, but could be objects extinguished a while ago, such as gamma-ray bursts \\citep[see, e.g.,][]{viet95,waxm95}, or sources spread over space like cosmological shocks \\citep[see, e.g.,][]{krj96,krb97}. The injection energy spectrum of power-law with cut-off at an arbitrary maximum energy (see Section 2.6) would be unrealistic. The IGMF in the Local Group (see Paper I), although currently little is known, might be strong enough to substantially deflect the trajectories of UHECRs. All of those will have effects on the quantitative results, which should be investigated further. (4) Recently, the Auger collaboration disclosed the analysis, which suggests a substantial fraction of highest energy UHECRs might be iron nuclei \\citep{auger07b,auger09d}. This is in contradiction with the analysis of the HiRes data, which indicates highest energy UHECRs would be mostly protons \\citep{st07}. The issue of composition still needs to be settled down among experiments. Iron nuclei, on the way from sources to us, suffer much larger deflection than protons. Hence, if a substantial fraction of UHECRs is iron, some of our findings will change, a question which should be investigated in the future." }, "0910/0910.3682_arXiv.txt": { "abstract": "Text{The giant planet atmospheres exhibit alternating prograde (eastward) and retrograde (westward) jets of different speeds and widths, with an equatorial jet that is prograde on Jupiter and Saturn and retrograde on Uranus and Neptune. The jets are variously thought to be driven by differential radiative heating of the upper atmosphere or by intrinsic heat fluxes emanating from the deep interior. But existing models cannot account for the different flow configurations on the giant planets in an energetically consistent manner. Here a three-dimensional general circulation model is used to show that the different flow configurations can be reproduced by mechanisms universal across the giant planets if differences in their radiative heating and intrinsic heat fluxes are taken into account. Whether the equatorial jet is prograde or retrograde depends on whether the deep intrinsic heat fluxes are strong enough that convection penetrates into the upper troposphere and generates strong equatorial Rossby waves there. Prograde equatorial jets result if convective Rossby wave generation is strong and low-latitude angular momentum flux divergence owing to baroclinic eddies generated off the equator is sufficiently weak (Jupiter and Saturn). Retrograde equatorial jets result if either convective Rossby wave generation is weak or absent (Uranus) or low-latitude angular momentum flux divergence owing to baroclinic eddies is sufficiently strong (Neptune). The different speeds and widths of the off-equatorial jets depend, among other factors, on the differential radiative heating of the atmosphere and the altitude of the jets, which are vertically sheared. The simulations have closed energy and angular momentum balances that are consistent with observations of the giant planets. They exhibit temperature structures closely resembling those observed, and make predictions about as-yet unobserved aspects of flow and temperature structures.} ", "introduction": "Among the most striking features of the giant planets are the alternating zonal jets. As shown in Fig.~\\ref{fig_flow_lat}, Jupiter and Saturn have prograde equatorial jets (superrotation) that peak at $\\about 100\\,\\mathrm{m \\, s^{-1}}$ and $\\about 200$--$400\\,\\mathrm{m \\, s^{-1}}$, depending on the vertical level considered. Uranus and Neptune have retrograde equatorial jets (subrotation) that peak at $\\about 100\\,\\mathrm{m \\, s^{-1}}$ and $\\about 150$--$400\\,\\mathrm{m \\, s^{-1}}$. Jupiter and Saturn have multiple off-equatorial jets in each hemisphere; Uranus and Neptune have only a single off-equatorial jet in each hemisphere. Despite decades of study with a variety of flow models, it has remained obscure how these different flow configurations come about \\citep{Vasavada05}. \\begin{figure*}[!htb] \\centering \\includegraphics{figure1} \\caption{Mean zonal velocities in the upper atmosphere of the giant planets from observations and simulations. Jupiter: observations from the \\emph{Cassini} spacecraft \\citep{Porco03} (orange line), and in simulation at $0.75$~bar (dark blue line). Saturn: observations from the \\emph{Voyager} spacecraft (orange line), from the \\emph{Hubble Space Telescope} (HST) \\citep{Sanchez-Lavega03} (green crosses), from the \\emph{Cassini} spacecraft at $\\about 0.06$~bar (magenta circles) and at $\\about 0.7$~bar (light blue squares) \\citep{Sanchez-Lavega07}, and in simulation at $0.1$~bar (dark blue line). Uranus: observations from the \\emph{Voyager} spacecraft (orange circles), HST (orange crosses) \\citep{Hammel01}, the Keck telescope (orange squares) \\citep{Hammel05}, and in simulation at $25.0$~mbar (dark blue line). Neptune: observations from the \\emph{Voyager} spacecraft (orange circles) and from HST (orange crosses) \\citep{Sromovsky01}, and in simulation at $25.0$~mbar (dark blue line). Differences between the statistically identical northern and southern hemispheres in the simulations are indicative of the sampling variability of the averages.} \\label{fig_flow_lat} \\end{figure*} Existing models posit as the driver of the flow either the differential radiative heating of the upper atmosphere \\cite[e.g.,][]{Williams79,Williams03a} or the intrinsic heat fluxes emanating from the deep interior \\citep[e.g.,][]{Busse76,Heimpel05,Aurnou07,Chan08,Kaspi09}. However, none of these models can account for the existence of equatorial superrotation on Jupiter and Saturn and equatorial subrotation on Uranus and Neptune with radiative heating, intrinsic heat fluxes, and other physical parameters consistent with observations. For example, deep-flow models that posit intrinsic heat fluxes as the sole driver of the flow can generate equatorial superrotation, but they use heat fluxes more than $10^6$ times larger than those observed \\citep[e.g.,][]{Heimpel07}. They generate equatorial subrotation only with intrinsic heat fluxes even stronger than those for which they generate superrotation \\citep{Aurnou07}, although the intrinsic heat fluxes on the subrotating planets (Uranus and Neptune) are weaker than those on the superrotating planets (Jupiter and Saturn). The relevance of such deep-flow models is further called into question by the eddy angular momentum fluxes they imply. Their meridional eddy fluxes of angular momentum per unit volume (taking density variations into account) have a barotropic structure: they extend roughly along cylinders concentric with the planet's spin axis over the entire depth of the fluid, typically to pressures of order $10^6$~bar \\citep[e.g.,][their Fig.~10]{Kaspi09}. But the eddy angular momentum fluxes inferred from tracking cloud features in Jupiter's and Saturn's upper tropospheres indicate that the mean conversion rate from eddy to mean-flow kinetic energy is of order $10^{-5} \\, \\mathrm{W \\ m^{-3}}$ \\citep{Ingersoll81,Salyk06,DelGenio07}. If the observed upper-tropospheric eddy fluxes of angular momentum per unit volume extended unabatedly over a layer of 50~km thickness (e.g., from about 0.3 to 2.5~bar pressure on Jupiter, or from about 0.3 to 0.9~bar pressure on Saturn), and if vertical zonal flow variations over this layer are not dramatic, the total energy conversion rate would be about $0.5\\,\\mathrm{W \\, m^{-2}}$. This is already $\\about 4\\%$ of the total energy uptake of the atmosphere from intrinsic heat fluxes and absorption of solar radiation for Jupiter, or $\\about 11\\%$ for Saturn. But the limited thermodynamic efficiency of atmospheres allows only a fraction of the total atmospheric energy uptake to be used to generate eddy kinetic energy \\citep{Lorenz55,Peixoto92}. The observations of Jupiter and Saturn therefore imply that eddy angular momentum fluxes cannot extend unabatedly over great depths and must have a baroclinic structure. Barotropic eddy angular momentum fluxes that extend to depths of order $10^6$~bar, with upper-atmospheric fluxes of similar scale and magnitude as those observed, are only possible in deep-flow models if the driving heat fluxes are several orders of magnitude greater than observed. Similarly, shallow-flow models that posit differential radiative heating as the driver of the flow can generate equatorial superrotation, but they require artifices such as additional equatorial heat or wave sources that have no clear physical interpretation \\citep[e.g.,][]{Williams03a, Williams03b, Yamazaki05, Lian08}. It is unclear in those and other shallow-flow models \\citep[e.g.,][]{Scott08} what physical characteristics distinguish the superrotating planets from the subrotating planets.\\footnote{\\citet{Lian10} claim that different rates of latent heat release in phase changes of water may be responsible for superrotation on Jupiter and Saturn and subrotation on Uranus and Neptune. However, they impose latent heat fluxes at the lower boundary of their model that are not consistent with the observed energetics of the planets. Similar to the simulations of \\citet{Aurnou07}, they require stronger energy (latent heat) fluxes to generate subrotation than to generate superrotation. For example, the latent heat fluxes are of order $10$--$20\\,\\mathrm{W\\,m^{-2}}$ in their Jupiter and Saturn simulations and of order $1500\\,\\mathrm{W\\,m^{-2}}$ in their Uranus/Neptune simulation (Y.~Lian, pers.\\ communication, 2010). The latter are several orders of magnitude larger than the observed intrinsic heat fluxes or absorbed radiative fluxes (Table~1), which would have to drive any latent heat fluxes (energy would be required to evaporate the condensate that falls from the upper atmosphere into deeper layers).} In \\citet{Schneider09} (SL09 hereafter), we postulated that prograde equatorial jets on the giant planets occur when intrinsic heat fluxes are strong enough that Rossby waves generated convectively in the equatorial region transport angular momentum toward the equator. Multiple off-equatorial jets, by contrast, form as a result of baroclinic instability owing to the differential radiative heating of the upper atmosphere. We introduced a general circulation model (GCM) and demonstrated with it that the postulated mechanisms can account qualitatively for large-scale flow structures observed on Jupiter. Here we use simulations with essentially the same GCM, with closed energy and angular momentum balances that are consistent with observations, to demonstrate universal formation mechanisms of jets on all the giant planets. We show that the different flow configurations on the giant planets can be explained through consideration of the different roles played by intrinsic heat fluxes and solar radiation in generating atmospheric waves and instabilities. Section~\\ref{s:GCM} briefly describes the GCM. Section~\\ref{s:Simulations} shows simulation results for Jupiter, Saturn, Uranus, and Neptune. Section~\\ref{s:Mechanisms} discusses the formation mechanisms of the jets in the simulations and confirms the postulated mechanisms through control simulations. Section~\\ref{s:AM_balance} discusses what the upper-atmospheric fluid dynamics, on which we focus, imply about flows at greater depth on the giant planets. Section~\\ref{s:Conclusion} summarizes the conclusions and their relevance for available and possible future observations. ", "conclusions": "\\label{s:Conclusion} We have presented the first simulations of all four giant planets with closed energy and angular momentum balances that are consistent with observations. The simulations reproduce many large-scale features of the observed flows, such as equatorial superrotation on Jupiter and Saturn and equatorial subrotation on Uranus and Neptune. They exhibit temperature structures that are broadly consistent with available observations, and they reproduce many details of the observed flows, for example, their vertical structure to the extent it is known and characteristic equatorial waves observed on Jupiter. We have demonstrated that equatorial superrotation is generated if convective Rossby wave generation is strong and low-latitude angular momentum flux divergence owing to baroclinic eddies generated off the equator is sufficiently weak (Jupiter and Saturn); equatorial subrotation results if either convective Rossby wave generation is weak or absent (Uranus) or low-latitude angular momentum flux divergence owing to baroclinic eddies is sufficiently strong (Neptune). Current computational resources limit our ability to simulate flows at depth. However, considerations of the angular momentum balance have shown that the zonal jets should extend---generally with shear---to the depth where drag acts on them and balances the angular momentum flux divergences and convergences in the upper troposphere. That drag acts on the zonal flow at depth is suggested by observations of eddy angular momentum fluxes on Jupiter and Saturn, and a plausible MHD drag mechanism exists. Though quantitative aspects (e.g., jet strength and shear) may be affected by our inability to resolve the flow and drag at depth, the jet formation mechanisms we discussed are not affected by it. We expect as-yet unobserved aspects of the flow and temperature structures to be consistent with the simulations and mechanisms we presented. For example, we predict that NASA's upcoming JUNO mission to Jupiter will find evidence of zonal flows with vertical shear similar to those in Fig.~\\ref{fig_circulation}: near the equator, a strong and deep prograde jet, and away from the equator, prograde jets that weaken and retrograde jets that weaken only slightly or strengthen with depth. The thermal and gravitational signature of such zonal flows will likely be measurable by JUNO. \\paragraph" }, "0910/0910.1364_arXiv.txt": { "abstract": "{} {We investigated the non-thermal hard X-ray emission in the Ophiuchus cluster of galaxies. Our aim is to characterise the physical properties of the non-thermal component and its interaction with the cosmic microwave background.} {We performed spatially resolved spectroscopy and imaging using XMM-Newton data to model the thermal emission. Combining this with INTEGRAL ISGRI data, we modelled the 0.6--140 keV band total emission in the central 7 arcmin region.} {The models that best describe both PN and ISGRI data contain a power-law component with a photon index in a range 2.2--2.5. This component produces $\\sim$10\\% of the total flux in the 1--10 keV band. The pressure of the non-thermal electrons is $\\sim$1\\% of that of the thermal electrons. Our results support the scenario whereby a relativistic electron population, which produces the recently detected radio mini-halo in Ophiuchus, also produces the hard X-rays via inverse compton scattering of the CMB photons. The best-fit models imply a differential momentum spectrum of the relativistic electrons with a slope of 3.4--4.0 and a magnetic field strength B=0.05--0.15 $\\mu$G. The lack of evidence for a recent major merger in the Ophiuchus centre allows the possibility that the relativistic electrons are produced by turbulence or hadronic collisions.} {} ", "introduction": "Non-thermal hard X-ray emission has been detected in several clusters of galaxies over the past few years (see Rephaeli et al., 2008 for a recent review). Since the detections have remained at the level of a few $\\sigma$, many models still remain as valid explanations for the phenomenon. The most popular explanation is the inverse compton scattering of the cosmic microwave background photons with the relativistic electrons in the cluster (e.g. Sarazin et al., 1988). In this model, the CMB photon ends up in the hard X-ray band since its energy increases by a factor of $\\sim 10^8$ via the scattering. In primary models, the relativistic electrons are originally thermal electrons that have been accelerated by a cluster merger (e.g. Sarazin \\& Lieu, 1998) or turbulence (e.g. Brunetti et al., 2001; Brunetti et al., 2004). In secondary models, the acceleration comes from hadronic collisions (e.g. Dennison, 1980; Pfrommer \\& Ensslin, 2004). Recently a radio mini-halo was detected in the centre of the Ophiuchus cluster with the VLA at 1.4 GHz (Govoni et al., 2009; Murgia et al., 2009). This proved the existence of a population of relativistic electrons in Ophiuchus. Ophiuchus is a hot (T$\\sim$9 keV) nearby (z=0.028) cluster located close to the Galactic centre (l $\\sim 1^{\\circ}$ , b $\\sim 9^{\\circ}$). Consistently, INTEGRAL detected excess emission over the thermal component at a 4--6$\\sigma$ level in the 20--80 keV energy band in the Ophiuchus cluster (Eckert et al., 2008). The analysis was lacking a sensitive instrument for modelling the thermal component at lower energies, but the excess was nevertheless consistent with having a non-thermal origin. Our aim in this work is to improve the modelling of the emission in the Ophiuchus centre utilising spatially resolved XMM-Newton spectroscopy. In this work we use the EPIC instruments PN for spectral analysis and MOS2 for imaging of the central 7 arcmin region of the Ophiuchus cluster. Because of the limitations of the spatial capability of INTEGRAL, we cannot exclude regions of complex temperature structure. Rather, we examine the spatially resolved XMM-Newton data in order to obtain an accurate model for the thermal emission to be used later with the INTEGRAL data. We also update the INTEGRAL analysis with additional data. \\begin{figure*}[tb] \\includegraphics[width=18cm,angle=0]{Oph_lc2.ps} \\vspace{-6cm} \\caption{The histogram shows the light curve of Ophiuchus in the full FOV in 1000 ks time bins in the hard band (E $>$ 10 keV for PN, left panel and E $>$ 9.5 keV for MOS2, right panel). The dotted horizontal lines show the adopted quiescent level. \\label{lc.fig}} \\end{figure*} ", "conclusions": "We analysed the central r=7 arcmin region of the Ophiuchus cluster of galaxies using data obtained with the XMM-Newton EPIC and INTEGRAL ISGRI instruments. The ISGRI data yielded a 5.7$\\sigma$ detection of excess emission in the 20--120 keV band over the thermal prediction, as modeled with PN data. Our XMM-Newton analysis confirmed the existence of a cool core in Ophiuchus. The derived very long cooling time (3 $\\times 10^{9}$ yr) as well as the lack of significant merger signatures argues against a recent major merger in the Ophiuchus centre. These features are consistent with most clusters hosting a radio mini-halo (Ferrari et al., 2008), which has also been detected in Ophiuchus (Govoni et al. 2009). In most proposed models for the non-thermal emission in clusters the same population of relativistic electrons produces both the radio emission (via synchrotron) and hard X-ray emission (via inverse compton scattering of the cosmic microwave background photons). Our results support this scenario, since the derived photon index of 2.2--2.5 for the non-thermal component corresponds to radio index of 1.2--1.5 with is consistent with the upper limit (1.7) derived from the radio observations of Ophiuchus (P\\'eres-Torrez et al., 2009). The non-thermal component produces $\\sim$10\\% of the total flux in the 1--10 keV band. These models imply a differential momentum spectrum of the relativistic electrons with a slope of 3.4-4.0 and a magnetic field strength B=0.05--0.15 $\\mu$G. The pressure of the non-thermal electrons is $\\sim$ 1\\% of that of the thermal electrons, i.e. the gravitational potential of the cluster is adequate for confining such a population of non-thermal electrons." }, "0910/0910.4451_arXiv.txt": { "abstract": "Massive neutron stars may harbor deconfined quark matter in their cores. I review some recent work on the microphysics and the phenomenology of compact stars with cores made of quark matter. This includes the equilibrium and stability of non-rotating and rapidly rotating stars, gravitational radiation from deformations in their quark cores, neutrino radiation and dichotomy of fast and slow cooling, and pulsar radio-timing anomalies. ", "introduction": "\\label{sec:Intro} Because of the quark substructure of nucleons predicted by quantum chromodynamics (QCD), nuclear matter will undergo a phase transition to quark matter if squeezed to sufficiently high densities. In the quark matter phase the ``liberated'' quarks occupy continuum states which, in the low-temperature and high-density regime, arrange themselves in a Fermi sphere. The Fermi-sphere determines the form of low-lying excitation spectrum of quark matter, which resembles that of less exotic low-temperature systems found in condensed matter (\\eg electron gas or ultracold atomic vapor) or hadronic physics (\\eg nuclear or neutron matter). In analogy to these system the attractive interaction between quarks, mediated by the gluon exchange (which is responsible for the bound state spectrum of QCD, \\eg, the nucleon and mesons) leads to quark superconductivity and superfluidity (color superconductivity) via the Bardeen-Cooper-Schrieffer mechanism~\\cite{Bailin:1983bm}. Furthermore, under stellar conditions the pairing between the two light flavors of quarks occurs at finite isospin chemical potential, \\ie, when the Fermi surfaces of up and down quarks are shifted apart by an amount which is of the same order of magnitude as the gap in the quasiparticle spectrum. Under these conditions the pairing between the fermions persists, but the actual pairing pattern may be significantly different from that of the BCS and is likely to involve breaking of the spatial symmetries by the condensate order parameter. During the last decade there has been a substantial progress in understanding of pairing in two-component asymmetric superconductors. Firstly, new developments on (variations of) the Larkin-Ovchinnikov-Fulde-Ferrell (LOFF) phase revealed novel lattice structures of the order parameter~\\cite{Bowers:2002xr}. Secondly, it has been suggested that such systems may actually phase separate into normal and superconducting domains~\\cite{Caldas} or the pairing may require changes in the shapes of the Fermi surfaces~\\cite{Muther:2002mc}. Controlled experiments on two-component cold atomic vapors with mismatched Fermi surfaces show that the phase separation scenario is at work. Whether this is the case in the related systems such as the isospin asymmetric nuclear matter or deconfined quark matter is still unclear. Therefore, a major goal of astrophysics of dense matter is to find and to quantify the manifestations of dense phases in the observable properties of compact stars. Here we review some recent progress towards this goal. ", "conclusions": "" }, "0910/0910.5727_arXiv.txt": { "abstract": "Late X-ray flares observed in X-ray afterglows of gamma-ray bursts (GRBs) suggest late central engine activities at a few minuets to hours after the burst. A few unambiguously confirmed cases of supernova associations with nearby long GRBs imply that an accompanying supernova-like component might be a common feature in all long GRB events. These motivate us to study the interactions of a late jet, responsible for a x-ray flare, with various components in a stellar explosion, responsible for a GRB. These components include a supernova shell-like ejecta, and a cocoon that was produced when the main jet producing the GRB itself was propagating through the progenitor star. We find that the interaction between the late jet and the supernova ejecta may produce a luminous (up to $10^{49}$ erg s$^{-1}$) thermal X-ray transient lasting for $\\sim 10$ s The interaction between the late jet and the cocoon produces synchrotron-self absorbed non-thermal emission, with the observed peak X-ray flux density from 0.001 $\\mu$Jy to 1 mJy at 1 keV and a peak optical flux density from 0.01 $\\mu$Jy to 0.1 Jy (for a redshift $z= 2$). The light curve due to the late jet - cocoon interaction has very small pulse-width-to-time ratio, $\\D t/t \\approx 0.01 - 0.5$, where $t$ is the pulse peak time since the burst trigger. Identifying these features in current and future observations would open a new frontier in the study of GRB progenitor stars. ", "introduction": "Long duration gamma-ray bursts (GRB) -- those lasting more than 2 s -- are thought to be produced by a relativistic outflow (or jet) with a kinetic energy $\\sim 10^{51} - 10^{52}$ erg (beaming effect corrected) when a massive star collapses at the end of its nuclear burning life cycle (see Piran 1999, 2005; M\\'{e}sz\\'{a}ros 2002 for reviews). The massive star origin of GRBs are supported by two different observations: (i) GRBs are found to be in actively star forming galaxies (e.g., Christensen et al. 2004, Castro Cer\\'on et al. 2006) or in star (especially massive ones) forming regions of the host galaxies (e.g., Paczy\\'{n}ski 1998; Bloom, Kulkarni \\& Djorgovski 2002; Fruchter et al. 2006); (ii) for a subset of nearly a dozen of GRBs, X-ray-rich GRBs and X-ray flashes, Supernova (SNa) features -- both temporally and spacially associated with the bursts (Woosley \\& Bloom 2006 for a review) -- were detected. For four of those: GRB 980425 (e.g., Galama et al. 1998), 030329 (e.g., Hjorth et al. 2003), 021211 (Della Valle et al. 2003) and 031203 (e.g., Malesani et al., 2004), the physically associated SNe were not only photometrically but also spectroscopically confirmed. The others of the subset show a late-time ($\\sim$ 10 days) SNa-like ``bump'' in the optical afterglows, with a simultaneous strong color evolution, e.g., in GRB 980326 (Bloom et al. 1999) and 011121 (Bloom et al. 2002), consistent with the hypothesis of an underlying SNa. The {\\it Swift} satellite has recently unveiled a ``canonical'' behavior pattern in about two-thirds of GRBs' early X-ray afterglows: a rapid decline phase lasting for $\\sim 10^2$ s is followed by a shallow decay phase lasting $\\sim 10^3 - 10^4$ s, then by a ``normal'' power-law decay phase and finally by a possible jet break (Nousek et al. 2006; O'Brien et al. 2006). In addition, X-ray flares are found in about 50\\% of all {\\it Swift} bursts; they have been discovered in all of the above phases (Burrows et al. 2005, 2007; Chincarini et al. 2007). Even long before {\\it Swift}, a late X-ray flare was detected by BeppoSAX for GRB 970508 (Piro et al. 1998). In this work we investigate a scenario in which a late jet -- responsible for producing the X-ray flares and possibly the shallow decay phase -- interacts with other components of a long-GRB stellar explosion. These components include a SNa ejecta, if a SNa explosion is accompanying the GRB event, and a cocoon created by the passage of the main GRB jet through the star (Ramirez-Ruiz et al. 2002, Matzner 2003, Zhang et al. 2004). In Section \\ref{sec:mul-com} we argue for the existence of the late jet and multiple components of a GRB explosion, and provide physical motivations of this work. We investigate the interactions of the late jet with the SNa ejecta in Section \\ref{sec:jet-ejecta} and with the cocoon in Section \\ref{sec:jet-cocoon} and calculate their associated emissions. The predicted emissions and their detection prospects are confronted with current observational data in Section \\ref{sec:obs}. The summary and implications are given in Section \\ref{sec:summary}. ", "conclusions": "\\label{sec:summary} Observations of X-ray flares in many GRB afterglows suggest the existence of a late jet from a long-lived central engine of a GRB at $\\sim 10^2$ s but possibly $10^4 - 10^5$ s after the main GRB event. Adopting the collapsing massive star origin for long-duration GRBs, and assuming that the supernova explosion to be at approximately the same time as the GRB, we have investigated the interactions of this late jet with the SNa ejecta and, with a cocoon that was left behind when the main GRB jet traversed the progenitor star. We find that late jet - SNa ejecta interaction should produce a thermal transient, lasting about $\\sim 10$ s and with a peak photon energy at a few keV, that should precede or accompany the flare. The luminosity of this transient is proportional to the kinetic luminosity of the late jet and can be as high as a few $\\times 10^{48}$ erg s$^{-1}$. The luminosity is smaller if the polar cavity created by the main GRB jet is only partially filled. This thermal transient is similar to the one associated with the breakout of the main GRB jet. Although it has a lower luminosity, its later occurrence makes it easier to detect. The observation of this signal can provide another evidence for the massive-star origin of GRBs and new information on the GRB - SNa association. The fact that no such thermal transient was observed so far implies that the late jet - SNa ejecta interaction is suppressed. This could happen if the polar region of the ejecta was evacuated by the main GRB jet and cavity has not sufficiently refilled itself, especially for a not-too-late jet, e.g., $t_F \\sim 10^2$ s, or the cavity could be kept open by a continuous, low-power jet. A strong thermal transient signal can also be blocked by the optically thick cocoon, which would be the case if the cocoon is slowly moving. An alternative possibility is of a failed supernova scenario in which the entire stellar envelope collapses in a free fall time of a few $\\times 10^2$ s. In this case there is no supernova associated with the GRB and the late jet, if it is late enough, does not have to cross the stellar envelope. The late jet interaction with the cocoon would cause a flare or rebrightening, superposed on the afterglow light curves, at both the optical and X-ray bands. This flare would have a pulse-width-to-time ratio $\\D t /t < 1$ (the expected distribution of $\\D t/t$ is similar to that for X-ray flares). Depending on model parameters, we find for a burst at a redshift $z=2$ that the peak flux density at optical $f_{\\nu_{opt}}$ ranges from 0.01 $\\mu$Jy to 0.1 Jy ($V$-band apparent magnitude 29 to 11.5) and at X-rays $f_{\\nu_X}$ ranges from 0.001 $\\mu$Jy to 1 mJy. For typical parameters $f_{\\nu_{opt}} \\sim 0.1$ mJy ($V$-band magnitude $\\sim 19$) and $f_{\\nu_X} \\sim 1$ $\\mu$Jy. Observational identification of this emission would verify the existence of the cocoon produced when the GRB jet traversed the progenitor star, thus it would be another confirmation of the collapsar model for long duration GRBs (e.g., MacFadyen \\& Woosley 1999; Ramirez-Ruiz et al. 2002; Matzner 2003; Zhang et al. 2004). The late jet - cocoon interaction might have already been detected in four GRB afterglows in which simultaneous X-ray and optical flares with $\\D t /t \\ll 1$ were observed after the prompt emission has died off (see Fig. \\ref{fig:obs-cand}). From those candidate events, one can learn about the energetics of late jet and the cocoon by utilizing the emission calculation presented in this paper. Let us consider the flare event in GRB 050904 as an example. We find the most probable model parameters -- for this burst at $z= 6.3$ and with $t_F= 70$ s -- that produce the observed peak $f_{\\nu_{opt}}$ and $f_{\\nu_X}$ (data from Bo\\\"{e}r et al. 2006; Cusumano et al. 2007; Gou, Fox \\& M\\'{e}sz\\'{a}ros 2007) to be $E_c \\approx 10^{51}$ erg, $\\G_c \\approx 20 - 50$, $E_j \\approx 10^{52}$ erg and $\\G_j \\approx 500$. Those high energetics seem consistent with the very luminous nature of both the burst and the flare. There are three cases in which no optical flare was detected at the time of a strong X-ray flare, even though a number of optical telescopes have been observing these bursts at the time of the X-ray flares (Fig. \\ref{fig:obs-non-cand}). This shows that not all X-ray flares are due to the late jet - cocoon interaction. However, neither a late jet nor the cocoon can be ruled out in these cases. It is possible that a low-power jet preceding the late jet with a total energy of at least $10^{49}$ erg had kept the cavity in the cocoon open, so that the late jet - cocoon interaction was suppressed. If correct this implies a low level continuous emission from the central engine at the level of $\\sim 10^{47}\\, (t/10 \\, {\\rm s})^{-2}$ erg s$^{-1}$ lasting for $\\sim 10^{2}$ s, assuming a radiation efficiency of $\\sim 0.1$. \\begin{figure} \\centerline{ \\includegraphics[width=6cm,angle=0]{Molinari_et_al_060418.ps} } \\centerline{ \\includegraphics[width=6cm,angle=0]{Molinari_et_al_060607a.ps} } \\centerline{ \\includegraphics[width=7cm,angle=0]{Rykoff_et_al_060904b.ps} } \\caption{The three GRBs that show prominent late X-ray flares but without simultaneous optical flare apparent in the afterglow light curve. {\\it Top}: GRB 060418; {\\it Middle}: 060607a (both from Molinari et al. 2007). {\\it Bottom}: GRB 060904b; blue triangles are BAT data extrapolated to X-ray band, magenta squares are XRT data and red circles are optical data (from Rykoff et al. 2009).} \\label{fig:obs-non-cand} \\end{figure}" }, "0910/0910.3953.txt": { "abstract": "The Standing Accretion Shock Instability (SASI) is commonly believed to be responsible for large amplitude dipolar oscillations of the stalled shock during core collapse, potentially leading to an asymmetric supernovae explosion. The degree of asymmetry depends on the amplitude of SASI, but the nonlinear saturation mechanism has never been elucidated. We investigate the role of parasitic instabilities as a possible cause of nonlinear SASI saturation. As the shock oscillations create both vorticity and entropy gradients, we show that both Kelvin-Helmholtz and Rayleigh-Taylor types of instabilities are able to grow on a SASI mode if its amplitude is large enough. We obtain simple estimates of their growth rates, taking into account the effects of advection and entropy stratification. In the context of the advective-acoustic cycle, we use numerical simulations to demonstrate how the acoustic feedback can be decreased if a parasitic instability distorts the advected structure. The amplitude of the shock deformation is estimated analytically in this scenario. When applied to the set up of \\cite{fernandez09a}, this saturation mechanism is able to explain the dramatic decrease of the SASI power when both the nuclear dissociation energy and the cooling rate are varied. Our results open new perspectives for anticipating the effect, on the SASI amplitude, of the physical ingredients involved in the modeling of the collapsing star. ", "introduction": "Despite decades of active research \\citep{colgate66,bethe85}, the core collapse supernovae mechanism remains elusive. The failure of the most sophisticated 1D models to explode the majority of massive progenitors \\citep{liebendorfer01} suggests that multidimensional effects are essential for a successful explosion. Understanding the hydrodynamical instabilities responsible for this symmetry breaking, and more specifically their nonlinear dynamics, is therefore required to understand the explosion mechanism. The region between the neutrinosphere and the shock deserves particular attention because two instabilities take place there: neutrino-driven convection \\citep{herant92,herant94,burrows95,janka96,foglizzo06}, and the newly discovered Standing Accretion Shock Instability (SASI) \\citep{blondin03, ohnishi06, foglizzo07, scheck08}. 2D simulations suggest that the complex fluid motions triggered by these instabilities could lead to a successful explosion either by helping the classical neutrino-driven mechanism \\citep{buras06b, marek09, murphy08b} or by a new mechanism based on the emission of acoustic waves from the proto-neutron star (\\cite{burrows06, burrows07a}, see however \\cite{weinberg08}). The large scale ($l=1-2$) induced asymmetry could also explain the high kick velocities of newly formed neutron stars \\citep{scheck04, scheck06} and may affect significantly their spin \\citep{blondin07a, yamasaki08}. The linear phase of the two instabilities has been described in details by \\cite{foglizzo06, blondin06, foglizzo07,yamasaki07,fernandez09a}. Neutrino-driven convective modes with a large angular scale can be stabilized by a fast advection of matter through the gain region, whereas SASI is always dominated by large scale modes. This linear argument favors SASI as the cause of the prominent $ l=1-2 $ shock oscillations observed in the simulations. However, a theoretical understanding of the nonlinear development and saturation of SASI is still missing. This was highlighted by the unexpected dramatic decrease of the SASI power observed by \\cite{fernandez09a} when both the nuclear dissociation energy and the cooling rate are varied. To shed more light on this issue, this paper proposes a predictive saturation mechanism for SASI. This is a first step toward understanding how the amplitude of SASI depends on the physical ingredients of the model (nuclear dissociation, equation of state, heating rate, rotation, magnetic fields). We propose that the saturation takes place when a parasitic instability (also called secondary instability) grows on the dominant SASI mode. These parasites feed upon its energy and destroy its coherence, leading to a turbulent flow and the saturation of the growing SASI mode. This saturation mechanism by parasitic instabilities is similar to the one proposed for the saturation of the MRI by \\cite{Goodman94, Pessah09a}. Two types of instabilities are considered as potential parasites for the SASI mode: the Kelvin-Helmholtz instability (hereafter KHi) grows near a maximum of vorticity, and the Rayleigh-Taylor instability (hereafter RTi) grows on negative entropy gradients. In either case the cause of the instability lies within the SASI mode, and therefore the growth rate of these parasites increases with the amplitude of the shock oscillations. The parasites are thus able to affect significantly the dynamics of the flow only if the amplitude of SASI oscillations exceeds a certain threshold. The main objective of this work is to estimate this threshold. Understanding the saturation mechanism of SASI requires in principle some understanding of the mechanism underlying its linear growth. The saturation scenario we propose is not restricted to a particular mechanism for the growth of SASI, but it can be understood more precisely in the framework of the advective-acoustic cycle \\citep{foglizzo02, blondin03, ohnishi06, foglizzo07, scheck08, fernandez09a, foglizzo09}. In this mechanism, a shock deformation creates an entropy-vorticity wave, whose downward advection generates an acoustic feedback. This acoustic wave further deforms the shock, thus closing the unstable cycle. If the coherence of the advected wave were broken by a parasitic instability before it created the acoustic feedback, this cycle would be stabilized and SASI saturated. In Sect.~2, we explain our method and approximations. Sections 3 and 4 are devoted to the linear study of the KHi and RTi in our setup. In Sect.~5 we study how the development of the parasites breaks the coherence of a SASI mode and decreases of the acoustic feedback. In Sect.~6 we apply our results to the setup of \\cite{fernandez09a} and compare our estimates with the results of their simulations. Finally, our results are discussed in Sect.~7 and summarized in Sect.~8. ", "conclusions": "\\subsection{SASI or neutrino-driven convection? } \\cite{fernandez09b} argued that the large amplitude $l=1$ oscillations appearing in their numerical simulations including iron dissociation and a heating function is due to neutrino-driven convection rather than SASI, since SASI is stabilized by iron dissociation according to \\cite{fernandez09a}. Does iron dissociation at the shock really prevent SASI from growing to large amplitudes in a realistic core collapse? In realistic simulations the compression factor at the shock can reach $\\sim 10 $, which corresponds to $ \\epsilon = 0.14 $ in the present study and a SASI amplitude of $6\\%$ of the shock distance $( r_{\\rm sh} - r_{\\rm *} ) $, quite smaller than without dissociation ($40\\%$). We point out however that a significant fraction of this amplitude decrease may be an artifact of the parameterization, which changed the cooling function by a factor 127 in order to keep the ratio $ r_{\\rm sh}/r_{\\rm *} $ constant. This parameterization has the great advantage of being insensitive to geometrical effects that may arise if the aspect ratio between the shock and the cooling surface is changed. However cooling is then artificially low when dissociation is taken into account without heating. As a consequence, the resulting flow profile may not be more realistic than the flow profile without dissociation. As our analysis suggests that entropy gradients play an important role in the saturation of SASI, we investigated the effect of keeping the cooling function constant when dissociation is varied, resulting in a change of the shock radius (from $2.5r_{*}$ to $1.46r_{*}$, for $0<\\epsilon < 0.2v_{\\rm ff}^{2}$). By performing the same analysis as in Section~6, we then find that the saturation amplitude of SASI should decrease significantly less than when the shock radius is kept constant : $\\Delta r/( r_{\\rm sh} - r_{\\rm *} ) $ decreases by a factor $1.75$ only. Equivalently, $\\Delta r/r_{\\rm sh} $ and $\\Delta r/r_{\\rm *} $ are decreased by a factor $3.5$ and a factor $6$ respectively (to be compared with a decrease of 15 when $r_{\\rm sh}$ is constant). Although the geometrical effects make a direct comparison difficult, the fact that all these numbers are significantly smaller than the former variation by a factor 15 suggests that the decrease of the cooling function, necessary to keep the shock radius constant, plays a key role in decreasing the saturation amplitude. More insight on this issue may be gained by including the effect of neutrino heating in a parameterized manner, such that dissociation could be varied while both the cooling function and the shock radius are constant. This calculation is left for a future study. The numerical simulations by \\cite{scheck08} suggest that SASI is able to grow to large amplitudes even in the presence of dissociation. These simulations are significantly more realistic than the set up studied here, since they include a realistic equation of state where dissociation is taken into account in a physical way, and a simplified treatment of neutrino heating and cooling. They also differed from those by \\cite{fernandez09a} by their choice of a moving inner boundary mimicking the proto neutron star contraction. In some of the models of this article (e.g. W00), neutrino-driven convection was artificially suppressed but still SASI oscillations could grow to non negligible amplitudes. %, contradicting the claim that dissociation prevents SASI from growing to large amplitudes. It is however difficult to determine which difference between the two setups affects most importantly the saturation amplitude of SASI. Incidentally, it is worth noting that RTi mushrooms have been identified growing on the SASI entropy gradients in Fig.~7 of \\cite{scheck08} and were interpreted in their Sect.~6.1 as secondary convection. Although in that article convection was not recognized as an agent of SASI saturation, the fact that the RTi appears at a SASI amplitude close to the saturation amplitude is consistent with a parasitic mechanism of saturation. The interaction between SASI and neutrino-driven convection is complex and still poorly understood. Could neutrino-driven convection prevent the growth of SASI by breaking its mode structure ? One may argue that neutrino-driven convection does not feed upon the SASI mode energy, but rather converts free energy from the stationary gradients into vorticity. Could neutrino-driven convection feed SASI, either by creating vorticity which would enter the advective-acoustic cycle, or by creating sound waves \\citep{fernandez09b}? These difficult questions are beyond the scope of our study. \\subsection{Distinguishing RTi from KHi in the simulations} The RTi is often characterized by finger-like or mushroom-like structures as in Fig.~\\ref{imageRT}, while the KHi is characterized by vortices as in Fig.~\\ref{imageKH} and \\ref{toyfigure_KH}. However, in a complex flow containing both entropy gradients, shear and advection, RTi mushrooms may look like vortices (Fig.~\\ref{toyfigure_RT}). The following criterion may be more useful: the RTi should occur preferentially where the entropy perturbation is maximum, while the KHi occurs where the shear is maximum. In a sloshing mode these two maxima are very distinct: the entropy oscillation is maximum at the pole (where the shock speed is maximum), while vorticity is maximum at the equator (where the inclination of the shock is maximum). The parasitic structures visible in the simulations by \\cite{scheck08} and in the movies published online by \\cite{fernandez09a} are more vigorous near the pole, in agreement with our analysis. Furthermore, the RTi structure should grow preferentially on the half wavelength of a SASI mode where the entropy gradient is negative. By contrast, the KHi should grow on the whole extent of the SASI wavelength. An inspection of Fig.~7 of \\cite{scheck08} and of the movies by \\cite{fernandez09a} confirms this distinct feature of the RTi. \\subsection{Numerical resolution needed to resolve the parasites} An interesting concern raised by this saturation mechanism is that simulations should be able to resolve the parasitic instabilities properly in order to give reliable results on the nonlinear behavior of SASI. The RTi is a short wavelength instability, but as is shown in Sect.~4.3 advection tends to stabilize the small scales and makes the RTi dominated by large scales. Entropy stratification on the other hand favors small scales. As in the set up studied here the RTi is found to develop where both stabilizing effects are important, it is hard to make any prediction for its dominant wavelength. Any convergence study should verify that the grid size allows for the growth of parasites. As an example, \\cite{scheck08} witnessed the growth of Rayleigh-Taylor mushrooms with a typical angular scale of $ l \\sim 20-30 $. They were able to capture this small angular scale by using $360$ angular zones for $180\\,^{\\circ} $. Most 2D simulations use a resolution with $>100-200$ angular zones, and would probably resolve these scales (200 zones in \\cite{murphy08b}, 121 in \\cite{burrows06}, 128-192 in \\cite{marek09}, and 60-120 zones in \\cite{ohnishi06}). However 3D simulations may not be able to resolve such small scales. For example \\cite{iwakami08, iwakami09} mostly use a resolution of $30$ angular zones for $180\\,^{\\circ}$, which may be too coarse to capture such a small scale behavior. We note that the saturation amplitude of the low-$l$ modes ($l=1-3$) in Fig.~16 of \\cite{iwakami08} is slightly smaller at ``high resolution\" (60 zones) than at ``low resolution\" (30 zones). While this may be explained by a suppression of parasitic instabilities at low resolution, we cannot exclude that a different saturation process may take place in 3D, as discussed below. \\subsection{Effects of other physical ingredients} The saturation mechanism described in this paper can be used to anticipate the effects of many other physical ingredients of the core collapse model (e.g. 3D versus 2D, the rotation rate, the magnetic field) although a detailed analysis is beyond the scope of this paper. \\begin{itemize} \\item \\emph{3D versus 2D}: 3D simulations allow for non-axisymmetric modes that are artificially forbidden in axisymmetric simulations, thus a greater number of modes is available to the linear development of SASI. A single mode $l=1$, $m=0$ often dominates in 2D, whereas $3$ modes $l=1$, $m=0,\\pm1$ have the same growth rate in 3D if the collapsing core does not rotate \\citep{foglizzo07}. Besides, \\cite{iwakami08} found that the saturated mode amplitude is independent of $m$. We cannot exclude that nonlinear processes associated with the coupling between different mode, ignored in our analysis, are more important in 3D than in 2D. Contrary to \\cite{iwakami08} however, \\cite{blondin07a} found that \\emph{one} spiral mode dominates the 3D dynamics. Assuming the parasitic growth of instabilities is the dominant saturation mechanism, our analysis based on a linear description of the parasites would predict the same saturation amplitudes of SASI in 2D or 3D. However the nonlinear behavior of the RTi is known to differ in 3D and 2D (e.g. \\cite{goncharov02}, \\cite{cabot06}), and this may affect the saturation of SASI. \\cite{iwakami08} reported a smaller saturation amplitude of each individual SASI mode in 3D as compared to 2D, although the numerical convergence of this result should be further checked (Section 7.3). If confirmed, it would raise the following questions : is this difference in amplitude a consequence of the different non linear Rayleigh-Taylor behavior in 3D? Or is this the signature of a different saturation mechanism based on the interaction of $m\\ne 0$ modes ? A more systematic parametric study, similar to \\cite{fernandez09a} but in 3D, could help check the relevance of parasitic instabilities in 3D. \\item \\emph{Rotation rate}: \\cite{yamasaki08} have shown that rotation increases the growth rate of the spiral modes rotating in the same direction as the steady flow, while stabilizing the counter-rotating ones. If the rotation is strong enough, a single spiral mode dominates the evolution of SASI \\citep{blondin07a, iwakami08}. According to our analysis (Eq.~(\\ref{satRT})), the larger growth rate of the spiral mode could lead to a larger saturation amplitude of SASI. Nevertheless, a detailed calculation using the exact entropy and vorticity profiles of the SASI eigenmodes in a rotating flow is required in order to make an accurate prediction. \\item \\emph{Magnetic field strength}: The effect of the magnetic field on the linear phase of SASI is yet to be understood \\citep{guilet10}, but its effect on parasitic instabilities can already be anticipated from the point of view of the magnetic tension which tends to prevent motions that distort the magnetic field lines. This effect is stabilizing for the perturbations with a wave vector parallel to the magnetic field, but does not affect those whose wave vector is perpendicular. One would then expect that the magnetic field does not change the maximum RTi growth rate, but selects RTi modes with a wave vector perpendicular to the field lines. In contrast, the KHi can be suppressed if the magnetic field along the direction of the transverse velocity is strong enough. In a situation where the KHi were the dominant parasitic instability, a magnetic field could potentially allow for a larger saturation amplitude. \\end{itemize}" }, "0910/0910.0875_arXiv.txt": { "abstract": "For more than 140 years the chemical composition of our Sun has been considered typical of solar-type stars. Our highly differential elemental abundance analysis of unprecedented accuracy ($\\sim0.01$ dex) of the Sun relative to solar twins, shows that the Sun has a peculiar chemical composition with a $\\approx 20$\\% depletion of refractory elements relative to the volatile elements in comparison with solar twins. The abundance differences correlate strongly with the condensation temperatures of the elements. A similar study of solar analogs from planet surveys shows that this peculiarity also holds in comparisons with solar analogs known to have close-in giant planets while the majority of solar analogs without detected giant planets show the solar abundance pattern. The peculiarities in the solar chemical composition can be explained as signatures of the formation of terrestrial planets like our own Earth. ", "introduction": "For many years people have wondered about how our Sun compares to other stars and to whether our existence is related to special properties of our solar system. Angelo Secchi compared the Sun to many bright stars (Secchi 1868). He classified the stars in three types: type I which is the modern class A and early F, type II which are M stars, and type III comprising modern class G, K and early F, and called by Secchi {\\it tipo solare} (solar type), due to their spectroscopic similarity to the Sun. Furthermore, he concluded that ``{\\it le stelle di questo terzo tipo mostrano di avere una composizione identica a quella del nostro Sole}'' (Secchi 1868), meaning that solar type stars have identical composition to our Sun. Further works (e.g. Payne 1925; Edvardsson et al. 1993; Reddy et al. 2003) have not conclusively shown whether the Sun has a normal composition or not, due to the relatively large ($\\gtrsim$ 0.05) remaining systematic errors (Gustafsson 2008). Thus, for more than 140 years father Secchi's conclusion on the ``universal'' solar composition of the Sun has remained valid. In order to make further progress, it is important to eliminate many of the systematic errors ($\\sim$ 0.05-0.1 dex) that plague stellar chemical composition analyses (Asplund 2005). Solar analogs, which are G0-G5 dwarfs, and solar twins, stars almost identical to the Sun (Cayrel de Strobel 1996), are important in this context, in particular solar twins, because due to their similarity to the Sun a highly differential analysis will cancel most systematic errors. Thus, the first step in accurate comparisons of the Sun to other stars is to find solar twins. After many years of intensive search (see review by Cayrel de Strobel 1996), Porto de Mello \\& da Silva (1997) found the first solar twin, 18 Sco [HD 146233], which is much closer to the Sun than previous candidates like 16 Cyg B [HD 186427] (see e.g. Fig. 2 of Mel\\'endez et al. 2006). More recent surveys have largely increased the number of solar twins in the field (Mel\\'endez et al. 2006; Mel\\'endez \\& Ram\\'{i}rez 2007; Takeda et al. 2007; Mel\\'endez et al. 2009; Ram\\'{i}rez et al. 2009) as well as in the open cluster M67 (Pasquini et al. 2008). The most productive survey at high resolution (R $\\sim$ 60,000 - 110,000) is being undertaken by our group. Most Northern solar twin candidates were observed with the 2.7m telescope at the McDonald observatory, and complemented with data obtained at the Keck observatory. The Southern targets were observed using the Magellan Clay 6.5 m telescope at Las Campanas, complemented with recent (August 2009) VLT observations. Our solar twin project started in 2002, when only one solar twin was known. In order to improve our chances of finding solar twins, we first expanded the color-temperature relations of Alonso et al. (1996) to other photometric systems (Mel\\'endez \\& Ram\\'{i}rez 2003), allowing us to use existing photometry in other systems (e.g. Geneva) to select the best targets. We later improved the calibrations adding more stars and including new homogeneous systems like Tycho ($B_T - V_T$) and 2MASS (Ram\\'{i}rez \\& Mel\\'endez 2005). Although our first solar twin proposal in 2004 did not fly, the same year a more ``exciting'' proposal on Li in halo stars was granted a few nights with Keck. A small amount of that observing time was devoted to twin candidates, resulting in the discovery of the second solar twin (HD 98618; Mel\\'endez et al. 2006) and revealing that our temperature scale (and that of Alonso et al. 1996), although precise, have probably a zero-point issue. Our new selection of candidates from the Hipparcos catalog took into account a preliminary zero-point offset, which has been recently confirmed by analyses of solar twins and used for improved temperature calibrations (Casagrande et al. 2009). New solar twin proposals at McDonald and Magellan at Las Campanas (through Australian access) during 2006 were successful. The first observing run at McDonald in April 2007 allowed us to identify the best solar twin known to date, HIP 56948 (Mel\\'endez \\& Ram\\'{i}rez 2007), which is not only very similar to the Sun physically, but has also a low Li abundance similar to solar. Its status of best solar twin has been recently confirmed by Takeda \\& Tajitsu (2009). On the other hand, the Magellan observations at Las Campanas are opening new windows for astrophysics of the 0.01 dex level in chemical abundance accuracy. New observations taken recently at the VLT with UVES and CRIRES, promise to achieve even better precision (0.005 dex, $\\sim$1\\%), and to use solar twins to set tight constraints on Li and Be depletion in the Sun (e.g. do Nascimento et al. 2009). \\begin{figure} \\begin{center} \\includegraphics[width=3.4in]{f01.eps} \\caption{Differences between [X/Fe] of the Sun and the mean values in the solar twins as a function of $T_{\\rm cond}$. The abundance pattern shows a break at $T_{\\rm cond}$ $\\sim 1200-1250$ K. The solid lines are fits to the abundance pattern, while the dashed lines represent the standard deviation from the fits. The low element-to-element scatter from the fits for the refractory ($\\sigma = 0.007$ dex) and volatile ($\\sigma = 0.011$ dex) elements confirms the high precision of our work. Observational errors (including errors in both the Sun and solar twins) are shown with dotted error bars, while the errors in the mean abundance of the solar twins are shown with solid error bars. } \\label{fig1} \\end{center} \\end{figure} ", "conclusions": "" }, "0910/0910.3667_arXiv.txt": { "abstract": "We compute the cross-correlation between a sample of 14,000 radio-loud AGN (RLAGN) with redshifts between 0.4 and 0.8 selected from the Sloan Digital Sky Survey and a reference sample of 1.2 million luminous red galaxies in the same redshift range. We quantify how the clustering of radio-loud AGN depends on host galaxy mass and on radio luminosity. Radio-loud AGN are clustered more strongly on all scales than control samples of radio-quiet galaxies with the same stellar masses and redshifts, but the differences are largest on scales less than $\\sim 1$~Mpc. In addition, the clustering amplitude of the RLAGN varies significantly with radio luminosity on scales less than $\\sim 1$~Mpc. This proves that the gaseous environment of a galaxy on the scale of its dark matter halo, plays a key role in determining not only the probability that a galaxy is radio-loud AGN, but also the total luminosity of the radio jet. Next, we compare the clustering of radio galaxies with that of radio-loud quasars in the same redshift range. Unified models predict that both types of active nuclei should cluster in the same way. Our data show that most RLAGN are clustered more strongly than radio-loud QSOs, even when the AGN and QSO samples are matched in both black hole mass and radio luminosity. Only the most extreme RLAGN and RLQSOs in our sample, with radio luminosities in excess of $\\sim10^{26}$~W~Hz$^{-1}$, have similar clustering properties. The majority of the strongly evolving RLAGN population at redshifts $\\sim 0.5$ are found in different environments to the quasars, and hence must be triggered by a different physical mechanism. ", "introduction": "In recent years, galaxy formation models have become increasingly interested in the radio AGN phenomenon, because it is hypothesized that these objects may regulate the star formation history and mass assembly of the most massive galaxies and black holes in the Universe. Nearby radio galaxies in clusters are observed to inject a significant amount of energy into the surrounding gas. As the radio jets expand and interact with the surrounding medium, they are believed to heat the gas and prevent further accretion onto the central galaxy. The precise conditions that determine whether an AGN develops radio jets/lobes are still a matter of debate. Several studies have shown that the probability for a galaxy to become radio-loud is a strong function of stellar mass and redshift (e.g. \\citealt{best05}; \\citealt{donoso}). The role that the environment plays in triggering or regulating the RLAGN phenomenon is not as well established. \\citet{ledlow} found that the fraction of radio sources and the shape of the bivariate radio-optical was the same for objects in cluster and field environments. \\citet{best07} found that group and cluster galaxies had similar radio properties to field galaxies, but the brightest galaxies at the centers of the groups where more likely to host radio-loud AGN than other galaxies of the same stellar mass. In the local universe, \\citet{mandelbaum} analyzed a large sample of RLAGN at z$\\sim$0.1. They showed that RLAGN inhabit massive dark matter halos ($>10^{12.5}$~M$_{\\odot}$) and that, at fixed stellar mass, radio-loud AGN are found in more massive dark matter halos than control galaxies of the same mass selected without regard to AGN properties. This result implies that RLAGN follow a different halo mass - stellar mass relation than normal galaxies. \\citet{mandelbaum} also found that the boost towards larger halo masses did not depend on radio luminosity. \\citet{hickox} investigated the clustering in a small sample of higher redshift radio-loud AGN selected from the AGN and Galaxy Evolution Survey (AGES). They found no difference in the clustering amplitude of radio galaxies when compared to normal galaxies matched in redshift, luminosity and color. Most nearby RLAGN lack any of the standard accretion-related signatures that would indicate that their black holes are growing significantly at the present day (\\citealt{hardcastle06}). In contrast, quasars are thought to be powered by supermassive black holes accreting at close to the Eddington rate. Large redshift surveys like the Two Degree Field Galaxy Redshift Survey (2dFGRS) and the Sloan Digital Sky Survey (SDSS) now provide angular positions, accurate photometry and spectra for tens of thousands of QSOs. Recent determinations of the quasar two-point correlation function have demonstrated that at $z<2.5$ quasars cluster like normal $L_*$ galaxies (\\citealt{croom}; \\citealt{coil}) and populate dark matter halos of $\\sim 10^{12}$~M$_{\\odot}$, with the clustering only weakly dependent on luminosity, color and virial black hole mass (\\citealt{shen1}). As one moves out in redshift, the radio-loud AGN population evolves very rapidly in radio luminosity. Whether the RLAGN population also evolves strongly in black hole accretion rate, is considerably less clear. In particular, our understanding of whether there is a relationship between powerful, high redshift radio-loud AGN and quasars is quite sketchy. Around 10\\% of the quasar population is radio-loud. Numerous investigations have found that radio-loud quasars and at least {\\em some} powerful radio galaxies share a number of common characteristics, such as excess infrared emission, comparable radio morphologies and luminosities, optical emission lines, large evolutionary rates, and host galaxies with similar properties. It has thus been tempting to link both phenomena under the hypothesis that they are the same active nuclei viewed at different orientations (e.g. \\citealt{barthel89}; \\citealt{urry}). A few facts are believed to be key in any attempt to understand the transition from the population of low-luminosity radio AGN produced by weakly accreting black holes at low redshifts, to a population of high-luminosity radio AGN that may be produced by strongly accreting black holes at high redshifts. \\citet{fanaroff} found an important correlation between radio morphology and radio power: low luminosity sources (Fanaroff-Riley Class I, FRI) show emission peaking close to the nuclei that fades toward the edges, whereas more luminous sources (Fanaroff-Riley Class II, FRII) are brightest toward the edges. \\citet{hine} discovered that radio galaxies could also be classified according to the strength of their optical emission lines: low excitation (weak-lined) radio galaxies or LERGs, and high excitation (strong-lined) objects or HERGs. Modern unification models usually associate quasars with the most powerful HERGs, and low luminosity LERGs with BL Lac objects. Although there is a notable correspondence between RLAGN luminosity, morphology and spectral type, i.e. lower luminosity FRIs with LERGs, and higher luminosity FRIIs with HERGs, the correlations between these properties are not straightforward. There are populations of FRI sources with high excitation nuclear lines, and conversely, FRII galaxies with low excitation spectra are also common. It has been known for years that very high redshift ($z>2$), powerful radio galaxies are often surrounded by galaxy overdensities with sizes of a few Mpc (e.g. \\citealt{pentericci}; \\citealt{miley}). Since we know that quasars at the same redshift are clustered like normal $L_*$ galaxies, this would seem to throw some doubt on a simple unified scheme for explaining both phenomena. In view of this highly complex situation, a more statistical approach to comparing the properties of quasars and radio galaxies may yield further insight. In this paper we present measurements of the projected cross-correlation between a sample of 14,000 radio-loud AGN with a median redshift of $z=0.55$ with the surrounding population of massive galaxies ($M_*>10^{11}$~M$_{\\odot}$). The large size of our samples allows us to investigate in detail how clustering depends on stellar mass and on radio luminosity. By comparing the RLAGN clustering with results from control samples matched in redshift, luminosity and mass, we isolate the effect that the radio AGN phenomenon has on the clustering signal. We cross-correlate radio quasars drawn from the SDSS with the same reference sample of massive galaxies. Again, by using control samples matched in black hole mass and radio luminosity, we ensure that we compare RLAGN and RLQSOs in as uniform a way as possible. This paper is organized as follows. In Section 2 we describe the surveys and samples used in this work. In Section 3 we explain the methodology adopted to calculate the two-point correlation function. Section 4 presents the results on radio-loud AGN and quasar clustering. Finally, in Section 5 we summarize our results and discuss the implications of this work. Throughout the paper we assume a flat $\\Lambda$CDM cosmology, with $\\Omega_{m}=0.3$ and $\\Omega_{\\Lambda}=0.7$. Unless otherwise stated, we adopt $h=H_{0}$/(100 km s$^{-1}$) and present the results in units of Mpc~$h^{-1}$ with $h=1$. ", "conclusions": "In this work, we have successfully applied cross-correlation techniques to characterize the environments of $\\sim 14,000$ radio-loud AGN with P$_{\\rm1.4 GHz}>10^{24}$~W~Hz$^{-1}$, selected from $\\sim 1.2$ million LRG at $0.41$~Mpc~$h^{-1}$ there is a weak, but significant anti-correlation with radio power. For $r_p<1$~Mpc~$h^{-1}$ the dependence of clustering amplitude on luminosity is more complex: the cross-correlation amplitude increases with luminosity up to $\\sim10^{25.3}$~W~Hz$^{-1}$, and then decreases for the most luminous radio sources in our sample. \\item We have compared the environments of radio-loud AGN and radio-loud QSOs. RLAGN are clustered more strongly than RLQSOs on all scales, indicate that they populate dark matter halos of different mass. These results hold even when the RLAGN and RLQSO samples are matched in radio luminosity and black hole mass. \\item There are indications that the very most luminous RLAGN and RLQSOs in our sample (P$>10^{26}$~W~Hz$^{-1}$) do have similar clustering amplitudes. Only at these very high radio powers are the space-densities of radio-loud quasars and radio galaxies similar. This implies that unification of the two AGN populations can only be valid above P$\\sim 10^{26}$~W~Hz$^{-1}$. \\end{itemize} One major limitation of this study with regard to constraining AGN unification scenarios, is that it is based purely on photometric data from the SDSS, so we are unable to split our RLAGN sample into high-excitation and low-excitation sources. It is quite possible that the presence or absence of emission lines will provide the best way to define a population of radio galaxies that are clearly unified with the quasars. In this case, we would expect to find that the high-excitation radio galaxy population would cluster in a similar way to the quasars. In addition, we note that because the parent sample of our RLAGN catalogue consists of luminous {\\em red} galaxies, it is also likely that we completely miss some number of RLAGN with bluer colors and stronger emission lines. The analysis of the RLAGN luminosity function presented in \\citet{donoso} indicates that the missing sources cannot constitute more than $\\sim$20\\% of the total RLAGN population, so will not dominate the clustering signal of the radio AGN population as a whole. Nevertheless the quasar analogues among the radio galaxy population may still be under-represented in our analysis. Fortunately, upcoming large spectroscopic surveys such as BOSS will target nearly complete samples of more than a million massive galaxies at $0.410^7$ stars. The Galaxy is an inherently complex object, and the task of interpreting observations is made yet more difficult by our location within it. Consequently, the ambitious goals that the community has set itself, of mapping the Galaxy's dark-matter content and unravelling how it was assembled, can probably only be attained by mapping observational data onto sophisticated models. We are developing a modelling strategy that has as its point of departure analytic approximations to the distribution functions (\\df s) of various components of the Galaxy (McMillan et al.\\ in preparation). In this paper we present such approximations for the thin and thick discs. The paper is organised as follows. Section 2 explains how the \\df\\ is assembled. Section 3 compares the \\df's predictions for various observables to data. In particular evidence is presented that the Sun's $V$ velocity is conventionally underestimated by $\\sim6\\kms$ and predictions are given for velocity distributions as a function of distance from the plane. Evidence is presented that the standard \\df\\ provides a cleaner division of solar-neighbourhood stars into members of the thin and thick discs than has been available hitherto. Section 4 sums up and looks ahead. ", "conclusions": "We have explored the ability of distribution functions to provide models of the thin and thick discs of the Milky Way. Our \\df s are analytic functions of the actions of orbits, which ensures that there is an intuitive relation between the observable properties of the population a \\df\\ describes and the functional form of the \\df, and a meaningful way to compare models that use different gravitational potentials. In this paper we have used expressions for the actions that are only approximate, and imply that a star's vertical motion is adiabatically invariant during the star's motion parallel to the plane. In a forthcoming paper (McMillan et al., in preparation) orbital tori will be used to eliminate this approximation, and thus quantify its validity. We have shown that the vertical density profile and kinematics of the disc are accurately modelled by the extremely simple \\df\\ (\\ref{powerdf}). However, we rejected this \\df\\ because it is essential to be able to break the \\df\\ for the thin disc down at least into contributions from stars of various ages, and ideally into contributions from ranges in both age and metallicity. That is, we must recognise that the Galaxy is built up of innumerable stellar populations of various ages and metallicities, and each population has its own \\df. In this paper we have only begun to explore the resulting complexity by ascribing a single \\df\\ to the thick disc and modelling the thin disc as a superposition of \\df s for stars of different ages. In reality both discs are chemically inhomogeneous and we should assign a distinct \\df\\ to the stars born at each time with each chemical composition \\citep[e.g.][]{SchoenrichBII}. Hence the \\df s presented in this paper should be considered building blocks from which more elaborate \\df s may be in due course constructed. Our most basic building block is a ``pseudo-isothermal'' population of stars. \\figref{fig:HF} shows that the vertical distributions of young stellar populations is well modelled by a pseudo-isothermal population. The density of a pseudo-isothermal population does not decline exponentially with $z$, but Figs.~\\ref{fig:thinD} and \\ref{fig:thinthick} show that, remarkably, the composite population produced by stochastic acceleration of stars does have an exponentially decreasing density profile. An excellent fit to the observed density profile of the entire disc is obtained when a pseudo-isothermal thick disc is added to the composite thin disc. The dispersion in $v_z$ of thin-disc stars increases from $17.4\\kms$ in the plane to $33\\kms$ at $2.5\\kpc$, while that of the thick-disc stars increases from $\\sim 35\\kms$ in the plane to $\\sim48\\kms$ at $2.5\\kpc$. The thick disc contributes to the solar cylinder 24 per cent of the luminosity contributed by the thin disc, or 19.4 per cent of the total luminosity of the disc. Even though we are assuming that the dynamical coupling between motions in and perpendicular to the plane is weak, two features of our \\df s lead to strong correlations between distributions in $v_R$ and $v_z$. One feature is the fact that random velocities must increase as one moves inwards, and the other is the simultaneous increases in $\\sigma_R$ and $\\sigma_z$ that are driven by stochastic acceleration of a coeval population. Comparison of Figs.~\\ref{fig:DF} and \\ref{fig:thinthick} show that, on account of this correlation, the distribution of local stars in the $(v_R,v_\\phi)$ plane is atypical of the stellar population of the whole solar cylinder in just such a way that our composite disc \\df\\ can simultaneously provide reasonable matches to the very different shapes of the distributions of GCS stars in $v_R$ and $v_\\phi$. The widths of the model distributions in $v_R$ and $v_\\phi$ are controlled by a single parameter, $\\sigma_{r0}$. The shape of the $v_R$ distribution is predetermined by our choice of the \\df s functional form. The value of the parameter $R_\\d$ provides limited control of the shape of the $v_\\phi$ distribution and we obtain the best fit to the observed distribution when this parameter is chosen such that the disc's surface density declines roughly exponentially with scale length $2.5\\kpc$, which happens to agree with the scale length inferred from near-IR star counts by \\cite{Robin03}. In principle the \\df\\ of the thick disc should be tightly constrained by the dependence on $z$ of the velocity dispersions $\\sigma_R$ and $\\sigma_z$. These dependencies have recently been determined for SDSS stars by two groups. Unfortunately, their results seem to be incompatible and the reasons for the conflict are unknown. The standard model provides an excellent fit to the seminal work of Kuijken \\& Gilmore, perhaps because the gravitational potential in which the \\df\\ is evaluated was partly fitted to that work. Some of the difficulties encountered here with fitting newer data may arise from inaccuracy of the potential used. A worthwhile exercise would be to fit data from the GCS, RAVE and SDSS surveys to models that combined \\df s of the type presented here with and a multi-parameter gravitational potential: by simultaneously fitting the parameters in both the \\df\\ and the potential, one should be able to obtain reasonable fits to the data, providing the latter have been purged of such evident inconsistencies as those seen in \\figref{fig:powersig}. Data from more than one survey would probably have to be used since SDSS stars are too faint to constrain the thin disc tightly, although the RAVE survey, which certainly probes the thick disc effectively, may include enough nearby stars to make the Hipparcos-based GCS survey obsolete. The model fit to the $v_\\phi$ distribution of GCS stars is far from perfect. Some of the disagreement arises because, as is well known, the Galactic bar and spiral arms give rise to features (``star streams'') in the local velocity distribution that are inconsistent with the Galaxy being axisymmetric and in a steady state, as our models assume. Our favoured model $v_\\phi$ distribution would fit the data better if the conventional value of the solar motion $V_\\odot$ were $\\sim6\\kms$ too low. Tentative support for such an increase in the $V_\\odot$ is provided by astrometry of stellar masers \\citep{Reid09,McMillanB09}, and any increase would also tend to bring the Galaxy more into line with the Tully--Fisher relation between $\\vc$ and $M_I$ for external galaxies. By systematically perturbing the velocities of all solar-neighbourhood stars, spiral structure might lead to the classical approach to the determination of $V_\\odot$ yielding an underestimate. Further work is required to explore this possibility, and at this stage we would merely stress that the systematic error in $V_\\odot$ is much larger than the formal errors given by DB98 and \\cite{AumerB}. In the models, the asymmetric drifts of both the thin and thick discs increase with height. A disc's asymmetric drift is largely controlled by its parameter $R_\\d$ and in the standard model the asymmetric drift of the thin disc exceeds that of the thick disc above $1\\kpc$ because we have adopted a slightly larger value of $R_\\d$ for the thick disc than for the thin disc. A popular strategy for assigning solar-neighbourhood stars to the thin or thick disc is to find the values taken by each disc's model \\df\\ at the star's location. The model \\df s used are perfectly ellipsoidal but we show that such \\df s provide poor approximations to the thick-disc component of the standard \\df, so a markedly cleaner separation of the two discs could be obtained by replacing the ellipsoidal \\df s by the thin- and thick-disc components of the standard \\df. Although the observational material relating to the Galaxy has increased enormously in recent years, we have shown that much of the available data can be successfully modelled with a simple analytical \\df. In a couple of aspects the data are in mild conflict with the \\df, but it is at least as likely that the fault lies with the data as the \\df. In the coming decade the volume and quality of the observational material available will increase dramatically. We anticipate that comparisons between each new data set and an evolving standard \\df\\ will reveal successes and failures similar to those encountered here. The successes will confirm the value of the \\df\\ as a summary of a large and inhomogeneous body of data, and the failures will lead to critical re-examination of both data and \\df. Sometimes the failure will arise from a defective calibration of the data or incorrect assumptions used in its reduction, and other times it will indicate that the \\df\\ is too simplistic. Either way we will learn something new and interesting. In this paper the \\df's parameters have been fitted to the data by eye and no attempt has been made to quantify uncertainties in parameter values. Clearly such uncertainties are important, and they could be most securely established by carrying the \\df's predictions closer to the raw observations than we have done. In future work probability distributions in colour--magnitude--proper-motion space, etc., should be predicted that can be compared with the actual star counts. Upcoming infrared surveys, such as the VHS with Vista and APOGEE, will probe the disc at remote locations. The predictions of the standard \\df\\ for those locations will be presented shortly, after orbital tori have been introduced as the means to convert between Cartesian and angle-action variables. This upgrade will make obsolete the approximation of adiabatically invariant vertical motions used here." }, "0910/0910.1724_arXiv.txt": { "abstract": "{} {We investigate the effect of the physical environment on water and ammonia abundances across the S140 photodissociation region (PDR) with an embedded outflow. } {We have used the Odin satellite to obtain strip maps of the ground-state rotational transitions of \\emph{ortho}-water and \\emph{ortho}-ammonia, as well as CO(5\\,--\\,4) and \\13co(5\\,--\\,4) across the PDR, and \\water18 in the central position. A physi-chemical inhomogeneous PDR model was used to compute the temperature and abundance distributions for water, ammonia and CO. A multi-zone escape probability method then calculated the level populations and intensity distributions. These results are compared to a homogeneous model computed with an enhanced version of the {{\\tt RADEX}} code. } {\\h2o, \\nh3 and \\13co show emission from an extended PDR with a narrow line width of $\\sim$3\\,\\kms. Like CO, the water line profile is dominated by outflow emission, however, mainly in the red wing. Even though CO shows strong self-absorption, no signs of self-absorption are seen in the water line. \\water18 is not detected. The PDR model suggests that the water emission mainly arises from the surfaces of optically thick, high density clumps with $n(\\mathrm{H_2}$)$\\ga$10$^{6}$\\,\\cmcub~and a clump water abundance, with respect to H$_2$, of 5\\x10$^{-8}$. The mean water abundance in the PDR is 5\\x10$^{-9}$, and between $\\sim$2\\x10$^{-8}$\\,--\\,2\\x10$^{-7}$ in the outflow derived from a simple two-level approximation. The {{\\tt RADEX}} model points to a somewhat higher average PDR water abundance of 1\\x10$^{-8}$. At low temperatures deep in the cloud the water emission is weaker, likely due to adsorption onto dust grains, while ammonia is still abundant. Ammonia is also observed in the extended clumpy PDR, likely from the same high density and warm clumps as water. The average ammonia abundance is about the same as for water: 4\\x10$^{-9}$ and 8\\x10$^{-9}$ given by the PDR model and {\\tt RADEX}, respectively. The differences between the models most likely arise due to uncertainties in density, beam-filling and volume filling of clumps. The similarity of water and ammonia PDR emission is also seen in the almost identical line profiles observed close to the bright rim. Around the central position, ammonia also shows some outflow emission although weaker than water in the red wing. Predictions of the \\h2o \\trans~and 1$_{1,\\,1}$\\,--\\,0$_{\\,0,\\,0}$ antenna temperatures across the PDR are estimated with our PDR model for the forthcoming observations with the Herschel Space Observatory. } {} ", "introduction": " ", "conclusions": "\\label{section conclusions} We have used the Odin satellite to observe water, ammonia and carbon-monoxide in the well-known molecular cloud S140. We have simultaneously observed five-point strip maps across the bright rim in S140 of the \\emph{ortho}-\\h2o \\trans~and the \\emph{ortho}-\\nh3 1$_0$\\,--\\,0$_{\\,0}$ transitions, as well as \\mbox{CO(5\\,--\\,4)} and \\mbox{\\13co(5\\,--\\,4)}. The \\nh3 transition has never previously been observed in S140. Observations of \\water18 in the central position resulted in a non-detection at a rms level of 8\\,mK. As support observation we also mapped \\mbox{\\13co(1\\,--\\,0)} with the Onsala 20m telescope. Like CO, the water line-profile is dominated by emission from a NW\\,--\\,SE outflow, however, mainly in the red wing. Strong self-absorption is seen in the optically thick CO emission, while no obvious signs are seen in the \\emph{ortho}-\\h2o, \\emph{ortho}-\\nh3 or the almost optically thin \\13co line profiles. In addition to the outflow, our water line shows emission from a more extended NE\\,--\\,SW elongated PDR. Both these components originate approximately around our central position. No additional emission closer to the bright rim or further into the molecular cloud is detected. Close to the bright rim the temperature is most likely too high for a detection of our transition with an upper state energy of 61\\,K. Instead, higher-lying transitions, observable with the Herschel Space Observatory, will have their peak intensity shifted towards the bright rim \\citep{2005A&A.440.559.Poelman.Spaans, 2006A&A.453.615.Poelman.Spaans}. Even closer to the bright rim at a few magnitudes of $A_\\mathrm{V}$, water is, however, photo-dissociated by the UV field. The \\emph{ortho}-\\nh3 emission seems to emanate from the same high density clumps in the PDR and the outflow as water, but also shows additional emission further into the cloud where the ambient gas temperature drops to about 30\\,K. Compared to water in the central position, ammonia has a weaker outflow emission in the red wing although similar in the blue outflow. Close to the bright rim, where the outflow contribution to the emission is very low, the water and ammonia line profiles are almost identical, suggesting an origin in the same gas and velocity fields of the PDR. The \\13co line also shows a very similar line profile as \\h2o and \\nh3 in this position with a narrow line width of $\\sim$3\\,\\kms. Abundances, with respect to H$_2$, in the PDR are estimated both with an enhanced version of the homogeneous {\\tt RADEX} code and with a clumpy PDR model. This model points to low mean water and ammonia abundances in the PDR of 5\\x10$^{-9}$ and 4\\x10$^{-9}$, respectively. In the high-density clumps both the average water and ammonia abundances increase to 5\\x10$^{-8}$. To match the observed PDR antenna temperatures with Odin and SWAS, a clumpy medium is required by the model with a high molecular hydrogen density in the clumps of $\\ga$1\\x10$^6$\\,\\cmcub. The resulting {\\tt RADEX} mean abundances are twice as high: 1.0\\x10$^{-8}$ and 8\\x10$^{-8}$ for water and ammonia, respectively, using a molecular hydrogen density of 4\\x10$^{5}$\\,\\cmcub~and a kinetic temperature of 55\\,K. The differences most likely arise from the uncertainty in density, beam-filling, and volume filling of the clumps. The opacity of the narrow PDR component of the \\trans~transition is constrained by the narrow line width, and is estimated by {\\tt RADEX} to be $\\sim$7. The PDR model also confirm a low water opacity with an unweighted mean opacity of 17 and a model range of $\\sim$10$^{-5}$\\,--\\,800. The mean outflow water abundance, derived from a simple two-level approximation, is higher than in the PDR by at least one order of magnitude, $\\sim$2\\x10$^{-8}$\\,--\\,2\\x10$^{-7}$. Predictions of antenna temperatures for observations with HIFI are given by our PDR model of the \\emph{ortho}- and \\emph{para}-\\h2o \\trans~and 1$_{1,1}$\\,--\\,0$_{0,0}$ transitions for nine positions across the bright rim, and are found to peak around 70\\,--\\,80\\arcsec~from the dissociation front in agreement with our observations." }, "0910/0910.2668.txt": { "abstract": "The slope and zero-point of the unevolved main sequence as a function of metallicity are investigated using a homogeneous catalog of nearby field stars with absolute magnitudes defined with revised $Hipparcos$ parallaxes, {\\it Tycho-2} photometry, and precise metallicities from high-dispersion spectroscopy. $(B-V)$ - temperature relations are derived from 1746 stars between [Fe/H] $= -0.5$ and +0.6 and 372 stars within 0.05 dex of solar abundance; for T$_e$ = 5770 K, the solar color is $B-V$ = 0.652 $\\pm$ 0.002 (s.e.m.). From over 500 cool dwarfs between [Fe/H] = $-0.5$ and +0.5, $\\Delta(B-V)/\\Delta$[Fe/H] at fixed $M_V$ = 0.213 $\\pm$ 0.005, with a very weak dependence upon the adopted main sequence slope with $B-V$ at a given [Fe/H]. At Hyades metallicity this translates into $\\Delta M_V/\\Delta$[Fe/H] at fixed $B-V$ = 0.98 $\\pm$ 0.02, midway between the range of values empirically derived from smaller and/or less homogenous samples and model isochrones. From field stars of similar metallicity, the Hyades ([Fe/H] = +0.13) with no reddening has $(m-M)_0$ = 3.33 $\\pm$ 0.02 and M67, with E$(B-V)$ = 0.041, $A_V$ = 3.1E$(B-V)$, and [Fe/H] = 0.00, has $(m-M)_0$ = 9.71 $\\pm$ 0.02 (s.e.m), where the errors quoted refer to internal errors alone. At the extreme end of the age and metallicity scale, with E$(B-V) = 0.125 \\pm 0.025$ and [Fe/H] $= +0.39 \\pm 0.06$, comparison of the fiducial relation for NGC 6791 to 19 field stars with $(B-V)$ above 0.90 and [Fe/H] = +0.25 or higher, adjusted to the metallicity of NGC 6791, leads to $(m-M)_0 = 13.07 \\pm 0.09$, internal and systematic errors included. ", "introduction": "Distance determination to open and globular clusters is key to placing them in the proper Galactic evolutionary context and an indispensable component in evaluating stellar evolution as a function of mass, chemical composition, and age. For nearby clusters like the Hyades, Praesepe, the Pleiades and Coma, $Hipparcos$ parallaxes \\citep{esa}, coupled with proper-motion and radial-velocity memberships, have generated precise distances, independent of the cluster's composition and reddening, though these results have not been without controversy \\citep{pi98, vl99, sn01, ma02, so05, vl09}. Any remote cluster beyond the reach of parallax with a known chemical makeup comparable to a nearby cluster can be compared differentially using stars on the unevolved main sequence to obtain a reliable distance (see, e.g., \\citet{pi04, an07}). While unevolved main sequences of nearby clusters are ideal reference points due to the uniformity of age and composition among the cluster stars, the range in chemical composition sampled by the nearby clusters is approximately [Fe/H]$ = -0.2$ to +0.2. For open clusters outside this range and for all globular clusters, a traditional fallback procedure is to use cooler field dwarfs with parallaxes and well-defined abundances either to isolate a stellar sample of similar abundance or to interpolate among stars that bracket the metallicity of interest \\citep{tw99, gr03}. Given the broad gaussian distribution in [Fe/H] for field stars, this approach becomes more of a challenge for clusters whose metallicity places them in the extended, low-metallicity tail of the field star sample. A routine alternative to field star comparison has been the construction of theoretical isochrones which can be tuned to any combination of composition and age, though ultimately these models must be linked to the empirical data of the field stars by matching the theoretical model combinations of mass, luminosity, composition and temperature to the observed values of mass, absolute magnitude, abundance and color at a given age. For stars of approximately solar mass and higher, there is reasonable agreement among the various isochrone compilations currently available in the literature and on-line, with most discrepancies among the models understandable in terms of differences in the adopted parameterizations of the internal physics, the model atmospheres, and the relations used to transfer from the theoretical to the observational plane \\citep{pad, bas04, de04, vrs06, ma08, des08}. As one might expect, the discrepancies within the observational color-magnitude diagram (CMD) plane grow larger as one moves to stars of lower mass and/or more extreme compositions, reflecting the same paucity of empirical constraints found when attempting a direct match of distant clusters to nearby field stars with parallaxes. Issues of reddening and metallicity determination aside, the more contentious discussions of cluster distances generally have focused on metal-poor systems of the thick disk and halo, while the metal-rich end of the distribution has been moderately immune due to the rich sample of nearby stars of solar abundance, the proximity of the Hyades, and the rarity of clusters with compositions significantly higher than the Hyades. For almost 50 years the one potential exception has been the old open cluster, NGC 6791, though the extent of its anomaly has been hidden beneath disagreements over its reddening, composition, distance, and age. In recent years significant progress has been made in constraining the first two parameters, both of which are critical to defining the third using main-sequence fitting, while the fourth has no bearing on the distance if the comparison is made to stars of sufficiently low mass. In every instance where revised reddening and/or metallicity estimates have been obtained, an improved distance modulus has been derived through comparison with theoretical isochrones \\citep{bed08, ca06, ca05, ki05, stet, ch99, kr95, tr95, ga94, mj94, de92}. For the apparent distance modulus, the current range from main sequence fitting extends from a low of 13.1 \\citep{stet} to a high of 13.6 \\citep{slv03,atm07}. While the scatter is partly due to the adoption of different values for [Fe/H] and E$(B-V)$, a large portion is tied to disagreement over the location of the unevolved main sequence among the isochrones at high metallicity and the range of the CMD used to define an adequate fit to the models; in some instances, the unevolved main sequence and giant branch features are used simultaneously to optimize the fit. To circumvent the issues presented by the discrepancies among metal-rich isochrones, it was decided that the distance to NGC 6791 could best be obtained by an empirical fit to unevolved field stars of similar composition. The feasibility of this option has been enhanced by the more restricted range among recent determinations of the cluster parameters, the availability of revised parallaxes and broad-band photometry from {\\it Hipparcos/Tycho-2} \\citep{esa,vl07}, and the compilation of a catalog of precise spectroscopic abundances, temperatures, and surface gravities on a common scale for almost 2100 nearby field stars. While the initial motivation for this study was the distance to NGC 6791, the need to define the precise location of the unevolved main sequence at high metallicity using the very restricted field star sample at comparable [Fe/H] generated a more comprehensive investigation of the metallicity dependence of the unevolved main sequence for stars of typical disk metallicity ([Fe/H] = -0.5 to +0.5). This paper builds upon the approach laid out in \\citet{pe03}, using a sample expanded by an order of magnitude to derive the change in $M_V$ for cool dwarf stars at a given $B-V$ as [Fe/H] is varied from -0.5 to +0.5. While it is usually assumed that the ratio, $\\Delta M_V/\\Delta$[Fe/H], is constant with metallicity for unevolved disk dwarfs, derived values from observation and theoretical models vary by a more than a factor of two \\citep{ko02, pe03, pi04, ka06}. The goal of this investigation is to explore the assumption of a constant ratio, derive it, and test the consistency of the field star main sequence when applied to well-studied open clusters. Unless noted otherwise, errors quoted are standard deviations. The layout of the paper is as follows. In Sec. 2 we will describe the database of astrometry, photometry, and spectroscopy for the nearby stars that is used in Sec. 3 to define the characteristics of the unevolved main sequence as a function of color and metallicity. In Sec. 4, the derived CMD relation is used to estimate distances to the well-studied open clusters, M67 and the Hyades, as well as the extremely metal-rich cluster, NGC 6791. Sec. 5 contains a summary of our conclusions. ", "conclusions": "Distance determination for field stars and clusters remains a primary observational objective for those interested in understanding stellar and Galactic evolution. Ideally, parallaxes for a large sample of nearby stars with well-defined abundances and temperatures/colors would allow one to map the impact on the absolute magnitude of varying the abundance of a star of a given temperature/color. With this information in hand, determining the distance to any cluster with reliable abundance and reddening information becomes a straightforward task. Unfortunately, the reality is somewhat different. While reliable parallaxes are available for a large sample of field stars and most have colors on the {\\it Tycho-2} system, the critical component missing from the picture has been a comparable catalog of precise metallicity estimates. The observational approaches to defining the ratio of $\\Delta M_V$ with [Fe/H] \\citep{ko02, pe03, ka06} have relied upon photometric abundances coupled to a spectroscopic subset of the sample. The slope of the relation is either assumed to be constant with [Fe/H] and color and/or tested through the use of theoretical isochrones. For \\citet{ko02} and \\citet{ka06}, the baseline used to define the slope is extended to [Fe/H] below $-1$ through the inclusion of halo dwarfs and/or globular clusters. The dataset of \\citet{ka06} extends to stars where evolution off the main sequence is significant and may produce a selection bias in the mean absolute magnitude with [Fe/H]. Starting with a dataset of almost 2000 stars with reliable parallaxes, spectroscopic abundances, and homogeneous colors, we have reduced the sample to approximately 500 stars between [Fe/H] = $-0.5$ and +0.5 with colors red enough that evolution off the main sequence should be negligible. However, as evidenced by the old, metal-rich turnoff stars in NGC 6791, even our metallicity-dependent cutoff allows some significantly evolved stars into the mix. Over the primary color range of interest, $B-V$ = 0.75 to 1.15, the ratio with metallicity, $\\Delta M_V/\\Delta$[Fe/H]$ = 0.98 \\pm$ 0.02 with no evidence for a color or [Fe/H] dependence. Because this value assumes a universal slope of 4.53 for the main sequence with $B-V$, it is probably better to define the relation in terms of $\\Delta(B-V)/\\Delta$[Fe/H]$ = 0.213 \\pm$ 0.005, which is only weakly dependent upon the adopted slope of the main sequence. The concern that the slope of the main sequence with $B-V$ varies with [Fe/H] is real and not dependent upon the multiple sets of theoretical isochrones which are inconsistent on this point. From the observed cluster sequences analyzed in this investigation, the Hyades has a main sequence slope of 4.5; over the same color range, the fiducial relation for NGC 6791 has a slope of 5.4. With the field-star relations in hand, reliable estimation of cluster distances should follow. The challenge in this phase is ensuring that the clusters are on the same photometric color and spectroscopic abundance system as the field stars which, for more distant objects, also includes accurate reddening estimation. Because our catalog includes Hyades stars and the Hyades is assumed to be reddening-free, the only potential source of controversy is the photometric scale. Our analysis confirms what has been found by others \\citep{jon06, an07, jon08}. The \\citet{joh55} photometric system in the Hyades is offset from the SAAO/E-region standards that define the transformed {\\it Tycho-2} system by approximately -0.009 mag and 0.02 mag in $B-V$ and $V$, respectively. When these corrections are applied to the fiducial relations for the Hyades \\citep{pi04}, the cluster produces an excellent match to the field stars at the appropriate [Fe/H] with the cluster modulus set at $(m-M)_0$ = 3.33. The second test of the system is a match to M67. Because this cluster and NGC 6791 are two steps further removed from the field star sample in that there are no cluster stars with spectroscopic abundances in our catalog and the photometric data are not directly comparable to the {\\it Tycho-2} system, we have applied the same color offset found for the Hyades to the cluster data and made differential comparisons based upon a commonly adopted reddening and metallicity estimate. If the M67 data of \\citet{sa04} are matched to the Hyades fiducial relation shifted to [Fe/H] = 0.00, the apparent modulus of the cluster becomes 9.81 $\\pm$ 0.02; if a match is made directly to the field stars of identical [Fe/H], the modulus increases by 0.03. The comparable value from \\citet{an07} and \\citet{pa08} is 9.765. Note that part of the offset relative to \\citet{an07} may be a product of the elegant but hybrid approach that defines the main sequence relations using a mix of theoretical isochrones with color corrections defined by cluster data. The isochrones used by \\citet{an07} define a ratio, $\\Delta M_V/\\Delta$[Fe/H], over the color range of interest that is typically 1.4, significantly larger than our derived value near 1. With the Hyades CMD position fixed via parallax, differential shifts with metallicity then define the predicted position of clusters like M67. Even for a change in [Fe/H] of 0.13, an overestimate of 25$\\%$ in the slope would lead to an underestimate of the distance by 0.03 mag. Finally, at the extreme end of the age and metallicity scales among open clusters, the distance to NGC 6791 is derived using field stars of comparable metallicity. Initial comparisons using the adjusted Hyades relation exhibited evidence for significant evolutionary effects off the main sequence for the bluer stars in our sample, implying that the average star in the comparison is older than the Hyades. This trend virtually disappears when the data are compared to the fiducial relation for the cluster. In fact, the bluest stars in the field are systematically fainter than the stars in the cluster at the same color, indicating that they are, as expected, younger than NGC 6791. Adopting [Fe/H] = 0.40 and E$(B-V)$ = 0.15, we find $(m-M)_0$ = 13.13 $\\pm$ 0.09. By comparison, the definitive value tied to analysis of the eclipsing binary, V20, using the same parameters is $(m-M)_0$ = 13.00 $\\pm$ 0.10. The significance of the difference is marginal, especially given the underlying question of the photometric zero-points in $V$ and $B-V$." }, "0910/0910.0980_arXiv.txt": { "abstract": "We consider a gravitational theory of a scalar field $\\phi$ with nonminimal derivative coupling to curvature. The coupling terms have the form $\\kappa_1 R\\phi_{,\\mu}\\phi^{,\\mu}$ and $\\kappa_2 R_{\\mu\\nu}\\phi^{,\\mu}\\phi^{,\\nu}$ where $\\kappa_1$ and $\\kappa_2$ are coupling parameters with dimensions of length-squared. In general, field equations of the theory contain third derivatives of $g_{\\mu\\nu}$ and $\\phi$. However, in the case $-2\\kappa_1=\\kappa_2\\equiv\\kappa$ the derivative coupling term reads $\\kappa G_{\\mu\\nu}\\phi^{,\\mu}\\phi^{,\\nu}$ and the order of corresponding field equations is reduced up to second one. Assuming $-2\\kappa_1=\\kappa_2$, we study the spatially-flat Friedman-Robertson-Walker model with a scale factor $a(t)$ and find new exact cosmological solutions. It is shown that properties of the model at early stages crucially depends on the sign of $\\kappa$. For negative $\\kappa$ the model has an initial cosmological singularity, i.e. $a(t)\\sim (t-t_i)^{2/3}$ in the limit $t\\to t_i$; and for positive $\\kappa$ the universe at early stages has the quasi-de Sitter behavior, i.e. $a(t)\\sim e^{Ht}$ in the limit $t\\to-\\infty$, where $H=(3\\sqrt{\\kappa})^{-1}$. The corresponding scalar field $\\phi$ is exponentially growing at $t\\to-\\infty$, i.e. $\\phi(t)\\sim e^{-t/\\sqrt{\\kappa}}$. At late stages the universe evolution does not depend on $\\kappa$ at all; namely, for any $\\kappa$ one has $a(t)\\sim t^{1/3}$ at $t\\to\\infty$. Summarizing, we conclude that a cosmological model with nonminimal derivative coupling of the form $\\kappa G_{\\mu\\nu}\\phi^{,\\mu}\\phi^{,\\nu}$ is able to explain in a unique manner both a quasi-de Sitter phase and an exit from it without any fine-tuned potential. ", "introduction": "} For many years scalar fields have been an object of great interest for physicists. The reasons for this are manifold. One of them is quite pragmatic: models with scalar fields are relatively simple, and therefore it appeared possible to study them in detail and then extrapolate the results to more realistic and complicated models. More physical motivations include such important topics as the idea about variable ``fundamental'' constants, the Jordan-Brans-Dicke theory initially elaborated to solve the Mach problem, the Kaluza-Klein compactification scheme, the low-energy limit of the superstring theory, and others. Scalar fields play an especially important role in cosmology. As a bright example, one may mention numerous inflationary models in which inflation in the early Universe is typically driven by a fundamental scalar field called an inflaton. Furthermore, a recent discovery of cosmic acceleration has only refreshed the interest to scalar fields which began to be considered as candidates to explain dark energy phenomena. The rather general form of action for a scalar-tensor theory of gravity with a single scalar field $\\phi$ can be given as\\footnote{Throughout this paper we use units such that $G=c=1$. The metric signature is $(- + + +)$ and the conventions for curvature tensors are $R^\\alpha_{\\beta\\gamma\\delta} = \\Gamma^\\alpha_{\\beta\\delta,\\gamma} - ...$ and $ R_{\\mu\\nu} = R^\\alpha_{\\mu\\alpha\\nu}$.} \\beq\\label{STTaction} S=\\int d^4x\\sqrt{-g}\\left\\{\\frac1{8\\pi}F(\\phi,R) -h(\\phi)g_{\\mu\\nu}\\phi^{,\\mu}\\phi^{,\\nu}\\right\\}, \\eeq where $g_{\\mu\\nu}$ is a metric, $g=\\det(g_{\\mu\\nu})$, and $R$ is the scalar curvature. Functions $F(\\phi,R)$ and $h(\\phi)$ are varying from theory to theory. The function $h(\\phi)$ is responsible for the sign of kinetic energy of the scalar field. For example, the choice $h(\\phi)\\equiv-1$ leads to a wide class of theories with the negative kinetic term. The function $F(\\phi,R)$, being in general nonlinear, provides a nonminimal coupling between a scalar field and curvature. Though a freedom in choosing of $F(\\phi,R)$ leads to an unlimited variety of scalar-tensor theories, it is known (see, for example, \\cite{Mae,FarGunNar,CroFra}) that there exist conformal transformations transforming this kind of theories to Einstein's theory with a new minimally coupled scalar field $\\phi$ and an effective potential $V(\\phi)$ describing its self-interaction. The potential $V(\\phi)$ is a very important ingredient of various cosmological models: a slowly varying potential behaves like an effective cosmological constat providing one or more than one inflationary phases. An appropriate choice of $V(\\phi)$ is known as a problem of fine tuning of the cosmological constant. A further extension of scalar-tensor theories can be represented by models with nonminimal couplings between derivatives of a scalar field and curvature. This kind of couplings may appear in some Kaluza-Klein theories \\cite{KK1,KK2} (see also \\cite{Lindebook}, Section 9.5). In 1993, Amendola \\cite{Ame} has been considered the most general gravity Lagrangian linear in the curvature scalar $R$, quadratic in $\\phi$, and containing terms with four derivatives including all of the following terms (see also \\cite{CapLamSch} for details): \\bea & L_1=\\kappa_1 R\\phi_{,\\mu}\\phi^{,\\mu};\\quad% L_2=\\kappa_2 R_{\\mu\\nu}\\phi^{,\\mu}\\phi^{,\\nu};\\quad% L_3=\\kappa_3 R \\phi\\square\\phi ; &\\nonumber\\\\ & L_4=\\kappa_4 R_{\\mu\\nu} \\phi\\phi^{;\\mu\\nu} ;\\quad% L_5=\\kappa_5 R_{;\\mu} \\phi\\phi^{,\\mu} ;\\quad% L_6=\\kappa_6 \\square R \\phi^2, &\\nonumber \\eea where coefficients $\\kappa_1,\\dots,\\kappa_6$ are coupling parameters with dimensions of length-squared. Using the divergencies $$ (R\\phi^{,\\mu}\\phi)_{;\\mu};\\quad (R^{\\mu\\nu}\\phi\\phi_{,\\mu})_{;\\nu};\\quad% (R^{;\\mu}\\phi^2)_{;\\mu}, $$ one may conclude that, without loss of generality, $L_4$, $L_5$, and $L_6$ are not necessary to be considered. Also one may rule out $L_3$ because it contains $\\phi$ itself, while coupling term of main interest are those, where only the gradient of $\\phi$ is included. Thus, a general scalar-tensor theory with nonminimal derivative couplings may include only two terms $L_1$ and $L_2$. As was shown by Amendola \\cite{Ame}, a theory with derivative couplings cannot be recasting into Einsteinian form by a conformal rescaling $\\tilde g_{\\mu\\nu}=e^{2\\omega}g_{\\mu\\nu}$. He also supposed that an effective cosmological constant, and then the inflationary phase can be recovered without considering any effective potential if a nonminimal derivative coupling is introduced. Amendola himself \\cite{Ame} has considered a cosmological model in the theory with the only derivative coupling term $L_1=\\kappa_1 R\\phi_{,\\mu}\\phi^{,\\mu}$ and, by using a generalized slow-rolling approximation (i.e., neglecting all terms of order higher than the second one), he has obtained some analytical inflationary solutions. A general model containing both $L_1$ and $L_2$ has been discussed in \\cite{CapLamSch} (see also \\cite{CapLam}); it was shown that the de Sitter spacetime is an attractor solution of the model if $4\\kappa_1+\\kappa_2>0$. Recently Daniel and Caldwell \\cite{DanCal} have considered a theory with the derivative coupling term $L_2=\\kappa_2 R_{\\mu\\nu}\\phi^{,\\mu}\\phi^{,\\nu}$; in particular, they studied constraints which precision tests of general relativity impose on the coupling parameter $\\kappa_2$. It is also worth mentioning a series of papers devoted to a nonminimal modification of the Einstein-Yang-Mills-Higgs theory \\cite{BalDehZay:07} (see also a review \\cite{BalDehZay} and references therein). In this paper we continue studying a scalar-tensor theory with nonminimal derivative couplings and construct new exact cosmological solutions of the theory. ", "conclusions": "We have considered the gravitational theory of a scalar field with nonminimal derivative coupling to curvature and studied cosmological models in this theory. The main results obtained are as follows: 1. The Lagrangian of the theory includes two derivative coupling terms $\\kappa_1 R\\phi_{,\\mu}\\phi^{,\\mu}$ and $\\kappa_2 R_{\\mu\\nu}\\phi^{,\\mu}\\phi^{,\\nu}$, where $\\kappa_1$ and $\\kappa_2$ are coupling parameters with dimensions of length-squared. In general, field equations of the theory are of third order, i.e., contain third derivatives of $g_{\\mu\\nu}$ and $\\phi$, but in the particular case the order of equations is reduced up to the second one. This case corresponds to the choice $-2\\kappa_1=\\kappa_2\\equiv\\kappa$, then a combination of derivative coupling terms turn into $\\kappa G_{\\mu\\nu}\\phi^{,\\mu}\\phi^{,\\nu}$. It is worth noting that Capozziello et al \\cite{CapLamSch} , at pages 43 and 47, have mentioned the case $-2\\kappa_1=\\kappa_2$ to play a special role, because it represents a singular point of the differential equation. In this paper, we have supposed that the theory with $-2\\kappa_1=\\kappa_2$ is more preferable with the physical point of view, since the corresponding field equations do not contain derivatives of dynamical variables of order higher than the second. 2. Assuming $-2\\kappa_1=\\kappa_2\\equiv\\kappa$, we have studied a cosmological model with the spatially-flat Friedman-Robertson-Walker metric. It was shown that a behavior of the scale factor $a(t)$ and the scalar field $\\phi$ at large times is the same for all values of $\\kappa$ including zero, that is the late evolution of universe does not depend on $\\kappa$. Namely, one has $a(t)\\sim t^{1/3}$ and $\\phi(t)\\sim\\ln t$ at $t\\to\\infty$. Note this asymptotical behavior coincides with that of the exact solution \\Ref{alpha0}, \\Ref{phi0} obtained for $\\kappa=0$ (no coupling). 3. General properties of the model crucially depends on a sign of $\\kappa$. For $\\kappa<0$ an asymptotical form of the cosmological metric for small times is given by Eq. \\Ref{metric_k<0}. A corresponding scale factor is $a(t)\\sim (t-t_i)^{2/3}$; it describes the universe with an initial singularity at $t=t_i$. A new interesting feature of the model with derivative coupling is that a behavior of the scalar field near the cosmological singularity is regular, $\\phi(t)\\sim t$ (see Eq. \\Ref{phi_k<0}). For $\\kappa>0$ the law of universe evolution is qualitatively distinct from that for $\\kappa<0$. Now at early stages the universe has the quasi-de Sitter behavior \\Ref{deSitter} corresponding to the cosmological constant $\\Lambda=(3\\kappa)^{-1}$. In the limit $t\\to-\\infty$ the scale factor has the following asymptotical form $a(t)\\sim \\exp\\left(\\frac{t}{3\\kappa^{1/2}}-\\frac{e^{2t/\\sqrt{\\kappa}}}{144\\pi\\kappa C^2}\\right)$ (see Eq. \\Ref{a_k>0}), hence $a(t)$ exponentially fast goes to the de-Sitter form $a(t)=e^{Ht}$ with $H=(3\\sqrt{\\kappa})^{-1}$. At the same time, the scalar field $\\phi$ is exponentially growing at $t\\to-\\infty$, namely $\\phi(t)\\sim e^{-t/\\sqrt{\\kappa}}$. In conclusion, let us summarize the most essential features of cosmology with nonminimal derivative coupling. First of all, we should emphasize that cosmological solutions with the quasi-de Sitter phase are typical solutions of the gravitational theory of a scalar field with derivative coupling of the form $\\kappa G_{\\mu\\nu}\\phi^{,\\mu}\\phi^{,\\nu}$ with positive $\\kappa$. So, in order to obtain an inflationary phase, one need no fine-tuned potential, and so one do not face with the problem of fine-tuning. Another important feature of the model consists in the fact that an exact cosmological solution with $\\kappa>0$ describes in a unique manner both a quasi-de Sitter phase and an exit from it. Thus, the problem of graceful exit from inflation in cosmology with the derivative coupling term $\\kappa G_{\\mu\\nu}\\phi^{,\\mu}\\phi^{,\\nu}$ has a natural solution without any fine-tuned potential. \\subsection*" }, "0910/0910.2396_arXiv.txt": { "abstract": "Recent surveys have revealed a lack of close-in planets around evolved stars more massive than 1.2~$M_{\\odot}$. Such planets are common around solar-mass stars. We have calculated the orbital evolution of planets around stars with a range of initial masses, and have shown how planetary orbits are affected by the evolution of the stars all the way to the tip of the Red Giant Branch (RGB). We find that tidal interaction can lead to the engulfment of close-in planets by evolved stars. The engulfment is more efficient for more-massive planets and less-massive stars. These results may explain the observed semi-major axis distribution of planets around evolved stars with masses larger than 1.5~$M_{\\odot}$. Our results also suggest that massive planets may form more efficiently around intermediate-mass stars. ", "introduction": "Observationally, due mostly to the inapplicability of high-precision Doppler techniques to stars with spectral types earlier than late-F (which have large rotational velocities and a small number of spectral lines), little was known until recently about the frequency of planets around the more-massive stars. This situation has now changed, as surveys have been extended to searches for planets around more-massive stars in an evolved stage \\citep{Doli07,Doli09,Frin02,Hat03,Hat05,Hat06,Joh07a,Joh07b,Joh08,Liu07,Lm07, Nied07,Nied09,Ref06,Sat07,Sat08,Set03,Set05}. One of the most important trends that these surveys have revealed is the lack of close-in planets orbiting stars with masses $M > 1.5~M_{\\sun}$ \\citep{Joh07b,Sat08,Wri09} despite the fact that these planets are found around $\\approx$20 \\% of the Main Sequence (MS) stars with $M < 1.2~M_{\\sun}$. The frequency of planets seems also to be higher around intermediate-mass stars \\citep{Lm07,Joh07b}. Furthermore, contrary to the correlation between metallicity and probability of planet-hosting found for solar-mass, main-sequence stars \\citep[e.g.,][]{Fv05,San05}, the more-massive planet-hosting stars do not exhibit higher metallicities \\citep[e.g.,][]{Pas07}. On the theoretical side, it has been shown that the formation of Jupiter-mass planets around M~stars may be hindered \\citep[e.g.,][]{L04,Il05}, while the probability that a given star has at least one gas giant increases linearly with the stellar mass up of 3~$M_{\\odot}$ \\citep[e.g.,][]{Kk08}. One possibility is that the observed difference in the orbital distribution of planets found around intermediate and solar-mass primaries is due to the evolution of the star. Since all of the planets orbiting red giants or subgiants with $M > 1.5~M_{\\sun}$ have semi-major axes $a > 0.5$~AU it has been suggested that the planets might be engulfed as the star evolves off the MS \\citep{Joh07b,Sat08}. This possibility is, however, often dismissed out in the literature with the argument that high-mass stars are physically too small to engulf hot Jupiters \\citep{Joh07b,Joh08,Cur09}. Another possibility is that the observed differences in orbital distribution are primordial, and and they are a consequence of the planet formation mechanism around more massive stars. Along these lines, it has been shown by \\cite{Cur09} that the dependence of the lifetime of the gaseous disk on the stellar mass could result in halting the inward migration of planets around high-mass stars, thus explaining the observed lack of short period planets around these stars. The point we are making in the present paper is that before stellar evolution can be ruled out as the mechanism behind the observed semi-major axis distribution of planets around evolved stars, detailed modeling of the orbital evolution needs to be performed. In other words, in order to determine the potential role of the stellar mass in the planet formation process, the effects of the evolution of the star on the observed orbit distribution around giants have to be correctly isolated. This is precisely the goal of this paper. ", "conclusions": "Our main goal has been to determine whether stellar evolution could explain the observed distribution of the semi-major axes of planetary orbits around evolved stars (i.e., semi-major axis $>$\\,0.5~AU). We found that when the details of the orbital evolution are accurately calculated, tidal interactions constitute a quite powerful mechanism, capable of capturing close-in planets into the envelope of evolved stars. To date, there are $\\sim$20 exoplanets discovered around giant stars with $M > 1.5~M_{\\sun}$. The host stars have radii in the range $0.02 < R_* < 0.1$~AU. Our models are consistent with the existence of these planets at the point in the RGB evolution at which they are observed (see e.g., Table~8 in \\citealt{Sat08} and our Figs.~\\ref{fig_2m} and~\\ref{fig_3m}). However, we do not expect to find massive planets with a $<$\\,0.4~AU around a 2~$M_\\sun$ star with a R$_*\\geq 0.1$~AU (or 24~$R_\\sun$). Our calculations provide the minimum orbital radius inside of which planets will be engulfed by the star at the end of the RGB evolution. We find that the evolution of the star alone can quantitatively explain the observed lack of close-in planets around evolved stars even allowing for the uncertainties associated with mechanisms such as mass loss along the RGB or tidal-interaction theory. A mechanism such as the one invoked by Currie (2009; i.e., a lifetime stellar-mass dependency of the gas in the planet-forming disk that can halt migration) is not needed, although it might still be present. We find that given an initial distance at which tidal capture is possible, the more massive the planet, the earlier it will be captured by the RGB envelope. Observationally, it appears that giant stars host more massive planets than MS stars \\citep[e.g.,][]{Joh07b,Lm07}. Since we find that more massive planets are expected to be engulfed earlier in the RGB evolution, the fact that they are more frequently observed may point towards a planet-formation mechanism that favors the formation of more-massive planets around intermediate-mass stars. This would be consistent with a scenario in which the disk mass scales with the stellar mass, and more massive disks produce more massive planets (see also \\citealt{Kk08}). Along similar lines, since we find a high probability of tidal capture of the planet by evolved stars, the higher frequency of planets observed around intermediate-mass stars \\citep{Lm07,Joh07b} seems to imply that the efficiency of planet formation must be considerably higher for more massive stars, compared to their solar analogous. \\cite{Ass09} estimated the probability of detecting transits of planetary companions to giant stars to be $\\geq$ 10 \\% for several of the known systems. Since tidal oribital decay decreases the initial orbital distance, it may increase the probability for the planet to be observed in transit. Our results suggest that many planets may be accreted by their host star \\citep[see also][]{SL99a,SL99b}. Although so far the results are based on a fairly limited sample, it appears that giant stars hosting planets have the same metallicity distribution as giant stars without planets \\citep{Pas07}. On the other hand, we should note that if a giant star engulfs a close-in planet early in the RGB, it will appear to be a giant star without a planet. This mechanism might perhaps provide for a partial explanation for the lack of correlation with metallicity. Finally, stars that have accreted planets may show a higher spin rate, as the planet's orbital angular momentum is transferred to the star \\citep[e.g.][]{LS02}. This phenomenon has been investigated by \\citet{Mass08} and \\citet{Carl09}, where they estimate the probability of finding rapid rotators among evolved stars. Our quantification of the tidal capture radius may help refine these calculations." }, "0910/0910.2675_arXiv.txt": { "abstract": "We report a moderate-depth (70 ksec), contiguous 0.7 \\sqdeg , \\chandra\\ survey, in the Lockman Hole Field of the {\\it Spitzer}/SWIRE Legacy Survey coincident with a completed, ultra-deep VLA survey with deep optical and near-infrared (NIR) imaging in-hand. The primary motivation is to distinguish starburst galaxies and Active Galactic Nuclei, including the significant, highly obscured (log \\nh $>$23) subset. \\chandra\\ has detected 775 X-ray sources to a limiting broad band (0.3-8 keV) flux $\\sim 4 \\times 10^{-16}$ \\fcgs . We present the X-ray catalog, fluxes, hardness ratios and multi-wavelength fluxes. The log N vs. log S agrees with those of previous surveys covering similar flux ranges. The Chandra and {\\it Spitzer} flux limits are well matched: 771 (99\\%) of the X-ray sources have IR (infared) or optical counterparts, and 333 have MIPS 24 $\\mu$m detections. There are 4 optical-only X-ray sources and 4 with no visible optical/IR counterpart. The very deep ($\\sim$2.7 $\\mu$Jy rms) VLA data yields 251 ($> 4 \\sigma$) radio counterparts, 44\\% of the X-ray sources in the field. We confirm that the tendency for lower X-ray flux sources to be harder is primarily due to absorption. As expected, there is no correlation between observed IR and X-ray flux. Optically bright, Type 1 and red AGN lie in distinct regions of the IR vs X-ray flux plots, demonstrating the wide range of SEDs in this sample and providing the potential for classification/source selection. Many optically-bright sources, which lie outside the AGN region in the optical vs X-ray plots ($\\rm f_r / f_x > 10$), lie inside the region predicted for red AGN in IR vs X-ray plots consistent with the presence of an active nucleus. More than 40\\% of the X-ray sources in the VLA field are radio-loud using the classical definition, \\rl . The majority of these are red and relatively faint in the optical so that the use of \\rl~to select those AGN with the strongest radio emission becomes questionable. Using the 24 $\\mu$m to radio flux ratio (\\q24 ) instead results in 13 of the 147 AGN with sufficient data being classified as radio-loud, in good agreement with the $\\sim$10\\% expected for broad-lined AGN based on optical surveys. We conclude that \\q24 ~is a more reliable indicator of radio-loudness. Use of \\rl ~should be confined to optically-selected, Type 1 AGN. ", "introduction": "SWIRE, the largest {\\it Spitzer} Legacy program, is tracing the evolution of dusty, star-forming galaxies, evolved stellar populations and AGN, as a function of environment from z \\gax 2.5 to the present epoch. SWIRE covers 6 fields with a total area of 49 \\sqdeg\\ in all seven {\\it Spitzer} bands, selected from the entire IRAS/DIRBE sky as those areas with the lowest 100$\\mu m$ surface brightness (which scales with $N_H$, Schlegel \\etal\\ 1998); the Lockman Hole is one of the two best regions, having several contiguous \\sqdeg\\ of low $N_H$($\\sim 7 \\times 10^{19}cm^{-2}$) sky. SWIRE's power comes from its large surface area, its depth, which probes the Universe out to redshifts z \\gax 2.5 and over time intervals $> 10 Gyr$, and its sensitivity to both evolved stellar populations (IRAC: 3--8$\\mu m$) and dusty objects (starbursts \\& obscured AGN; MIPS: 24,70,160$\\mu m$). Thus their distribution relative to environment, and evolution with respect to time and to the development of structure, can be studied together. SWIRE is one of several, on-going, medium-depth surveys which fill the gap between the deep and shallow surveys, dramatically illustrating the need for a multi-layered approach to efficiently fill the L-z plane in several wave-bands. The SWIRE prime science goal is to study the structure, evolution and environments of AGN, starbursts and spheroidal galaxies out to z\\gax 2.5. Key to this extensive multi-wavelength campaign is an X-ray survey deep enough to distinguish starbursts and AGN, including the significant numbers which are highly obscured \\nh\\gax 10$^{23}$ cm$^{-2}$. We carried out a \\chandra\\ X-ray survey in the best (in terms of Galactic extinction and absence of nearby bright sources, including no bright radio sources) extragalactic $\\sim$1 \\sqdeg\\ field within SWIRE. The X-ray observations cover 0.7 \\sqdeg\\ contiguously. The broad band (0.3-8.0 keV) flux limit is $4 \\times 10^{-16}$ \\fcgs\\ at the field centers, sufficient to detect all SWIRE IR AGN except for those with high absorption at low redshift (z$\\leq$0.8, log N$_H \\geq$ 24) while IR galaxies are not detected. Thus we are able to distinguish all but the most highly obscured AGN from amongst the IR sources by their X-ray emission. The standard AGN model includes a super-massive black hole surrounded by an accretion disk in the center of a galaxy. The central regions of the AGN produce strong, hard X-ray emission (and relativistic jets in radio-loud sources) along with thermal emission from an accretion disk. Gas in the vicinity is heated by the nuclear emission and produces the broad and narrow emission lines characteristic of Type 1 and Type 2 AGN respectively. The viewing orientation of an AGN effects its classification, as demonstrated by the detection of polarized broad lines in NGC1068 (Antonucci \\& Miller 1985). This result led to a general acceptance that some fraction of Type 2 AGN are edge-on Type 1s. Such unification schemes require highly opaque dust surrounding the accretion disk. In addition to obscuring our view to the nucleus, this dust is heated by the nuclear source producing strong IR emission. Thus the observed Spectral Energy Distribution (SED) of an AGN is dependent on viewing angle and the traditional optical/ultraviolet surveys are incomplete to obscured sources ({\\it e.g.} Polletta \\etal\\ 2006). It is now clear that there is a significant obscured AGN population which has been missed from earlier surveys. The powerful combination of radio, X-ray and IR observations available for the current surveys facilitates a different and potentially unbiased view of the AGN in the field and so probes the full AGN population, including obscured sources. The variety of new techniques has resulted in new types of AGN being found, including: red AGN (Cutri \\etal\\ 2002), X-ray bright optically-normal galaxies (XBONGS, Comastri et al. 2002), Type 2 quasars (Norman et al. 2002), X-ray detected Extremely Red Objects (V/EROs, Alexander \\etal\\ 2002, Brusa \\etal\\ 2005). Despite the several multi-wavelength surveys in progress during this era of Great Observatories, a full view of the AGN population remains elusive. Shallow surveys, which are dominated by Type 1 AGN, find largely distinct subsets of the population depending on waveband (Hickox et al. 2009). While overlap is significant in deeper surveys, selection in any single waveband does not find all the AGN (Barmby et al 2006, Polletta et al. 2006, Donley et al. 2008, Park et al. 2008). Samples of obscured AGN remain relatively small and biases are still in the process of being understood, so that the nature of the new AGN and their significance to the population as a whole remain undetermined. Compton thick AGN are hard to find because their X-ray flux is obscured to energies \\gax 10 keV. Estimates of the fraction of the general population which are Compton thick vary. X-ray selected AGN include $\\sim 30$\\% (Polletta \\etal\\ 2006, Treister et al. 2004) Compton-thick sources. This number doubles to $\\sim 66\\%$ when IR-selected AGN are included (Polletta et al. 2006, Treister et al. 2009). Estimates generated by stacking X-rays from IR-selected AGN also suggest a factor of two increase in IR-selected compared with X-ray selected Compton thick AGN (Fiore et al. 2008, Daddi et al. 2007), although the absolute numbers differ by factors of $\\sim$100 depending on selection techniques and assumptions. The latest estimate from modeling the Cosmic X-ray Background (CXRB) postulates an obscured population larger than the unobscured by a factor of 8 at low luminosities, half of which are Compton thick. This decreases to a factor of 2 at high luminosities (Gilli, Comastri \\& Hasinger 2007). Alternatively, many X-ray-selected, low-luminosity AGN may be intrinsically X-ray hard rather than obscured, possibly reducing the fraction of obscured (including Compton thick) AGN to $\\sim$20\\% (Hopkins et al. 2009). A number of wide area, multi-wavelength surveys with sufficient depth to view the full AGN population are currently in progress, including: ECDFS (Lehmer et al. 2005, 0.3 deg$^2$), OPTX (Trouille et al. 2008, 1 deg$^2$ non-contiguous), SWIRE/\\chandra\\ (this paper, 0.7 deg$^2$), AEGIS-X (Laird et al. 2009, 0.67 deg$^2$), and C-COSMOS (Elvis et al. 2009, 0.5+0.4 deg$^2$). They are beginning to provide significant samples of the relatively rare sources, such as the high redshift or obscured AGN, found in small numbers in the deep \\chandra\\ surveys (Luo et al. 2008, Alexander et al. 2003). These various surveys will allow us to properly characterize these populations and understand their significance to AGN as a whole. In this paper we present the 0.7 \\sqdeg~SWIRE/\\chandra\\ X-ray source catalog, SWIRE identifications and IR, optical and radio fluxes. IR properties of X-ray sources range from the power-law shape characteristic of AGN-dominated sources, mostly unobscured Type 1 AGN, to SEDs which are dominated by star formation or host-galaxy emission, implying obscuration of the AGN at these wavelengths (Franceschini et al. 2005, Barmby et al. 2006, Polletta et al. 2007, Feruglio et al. 2008, Cardamone et al. 2008, Gorjian et al. 2008). X-ray hardness is loosely related to the optical/IR colors, generally supporting this view. We characterize the X-ray and multi-wavelength properties of the SWIRE/\\chandra\\ sample and take an initial look at these properties as a function of X-ray hardness and radio-loudness. Detailed study and modeling of the spectral energy distributions (SEDs) will be presented in a companion paper (Polletta \\etal\\, in prep.). ", "conclusions": "We present a list of 775 \\chandra\\ X-ray sources in the SWIRE/\\chandra\\ medium-depth, X-ray survey. Cross-correlation with {\\it Spitzer}, optical and radio images of part/all of the field resulted in 771 (99\\%) identifications in at least one optical/IR band, 767 have IR counterparts visible in the {\\it Spitzer} data in at least one (IRAC) band and 333 have 24$\\mu$m with MIPS detections. Four of the sources have no optical/IR counterpart down to our flux limits and 4 have an optical but no IR couterpart. We present multi-wavelength flux measurements for 744 X-ray sources, all those which are uncontaminated, unconfused and above the formal survey thresholds. The near-IR$-$X-ray datasets are well-matched in flux limit and go deep into the AGN population, providing an excellent dataset for multi-wavelength studies of the full AGN population, which will be reported in a companion paper (Polletta et al., in prep.). As in earlier surveys (DK04), there is no correlation between X-ray hardness and hard X-ray flux in this sample, confirming that hardness is predominantly caused by obscuration in the X-rays. The very deep (2.7 $\\mu$Jy at the field center) VLA data, covering part of the \\chandra\\ field (Figure~\\ref{fg:image}), results in 251 ($> 4 \\sigma$) radio detections, $44$\\% of the 568 X-ray sources in the VLA field. Even the very deepest radio data cannot detect all the X-ray sources. We demonstrate that the traditional radio-to-optical flux ratio, \\rl, used to define radio-loudness in AGN breaks down for a large proportion of the sources in this X-ray selected sample due to the weakness of the optical emission. Use of the 24$\\mu$m flux in place of the optical, the \\q24 ~parameter (Appleton et al. 2004), brings the radio-loud fraction down to expected levels (9\\%) and is strongly preferred as an indicator of radio-loudness for the full AGN population. The wide range of optical/IR/X-ray SEDs in this X-ray selected sample is demonstrated by the lack of any correlation between the optical or IR flux and the X-ray flux. Comparison with tracks for optically-selected, Type 1, and red AGN SEDs demonstrates the predominance of red AGN in this sample. There is a continuous distribution rather than distinct classes of AGN, and the IR vs X-ray plots allow better discrimination than the optical as the effects of obscuration are much lower. Comparison of the source properties (X-ray hardness, flux ratios and radio loudness) with the predictions of the standard median SEDs allows us to broadly classify the source types as a function of position in flux-flux plots. A correlation between the 24$\\mu$m flux and the X-ray hardness disappears when the hard X-ray flux is used, demonstrating that it is due to absorption in the X-rays while the 24$\\mu$m flux remains uneffected. This reinforces the use of \\q24 ~to define radio-loudness due to its stability." }, "0910/0910.5325_arXiv.txt": { "abstract": "% During the period July 2007 - January 2009, the AGILE satellite, together with several other space- and ground-based observatories monitored the activity of the flat-spectrum radio quasar 3C~454.3, yielding the longest multiwavelength coverage of this \\gray quasar so far. The source underwent an unprecedented period of very high activity above 100 MeV, a few times reaching \\gray flux levels on a day time scale higher than $F=400 \\times 10^{-8}$\\,\\phcmsec, in conjunction with an extremely variable behavior in the optical $R$-band, even of the order of several tenth of magnitude in few hours, as shown by the GASP-WEBT light curves. We present the results of this long term multiwavelength monitoring campaign, with particular emphasis on the study of possible lags between the different wavebands, and the results of the modeling of simultaneous spectral energy distributions at different levels of activity. ", "introduction": "% Among active galactic nuclei (AGNs), blazars show intense and variable \\gray emission above 100~MeV \\citep{Hartman1999:3eg}, and the variability time scale can be as short as a few days, or last a few weeks. They emit across several decades of energy, from the radio to the TeV energy band and their spectral energy distributions (SEDs) are typically double humped with a first peak occurring in the IR/optical band in the flat-spectrum radio quasars (FSRQs) and low-energy peaked BL Lacs (LBLs), and at UV/X-rays in the high-energy peaked BL Lacs (HBLs). This peak is commonly interpreted as synchrotron radiation from high-energy electrons in a relativistic jet. The second SED component is commonly interpreted as inverse Compton (IC) scattering of soft seed photons by relativistic electrons, and peaks in the MeV--GeV and in the TeV energy bands in the FSRQs/LBLs and in the HBLs, respectively. A recent review of the blazar emission mechanisms and energetics is given in \\cite{Celotti2008:blazar:jet}. The FSRQ \\source{} (PKS~2251$+$158; $z=0.859$) is certainly one of the most active extragalatic sources at high energy. In the \\egret{} era, it was detected in 1992 during an intense \\gray flaring episode \\citep{Hartman1992:3C454iauc, Hartman1993:3C454_EGRET} when its flux $F_{\\rm E>100MeV}$ was observed to vary within the range $(0.4-1.4) \\times 10^{-6}$\\,photons\\,cm$^{-2}$\\,s$^{-1}$. In 1995, a 2-week campaign detected a \\gray flux $< 1/5$ of its historical maximum \\citep{Aller1997:3C454_EGRET}. Figure~\\ref{Fig:3c454:egret} shows the \\gray light curve for $E>100$\\,MeV as observed by EGRET in the period 1991--1995. \\begin{figure}[!ht] \\begin{center} \\includegraphics[width=8cm]{vercellones_f1.eps} \\end{center} \\vspace{-0.5truecm} \\caption{EGRET \\source{} light curve for $E>100$\\,MeV in the period 1991-1995. The downward arrow represents a $2\\sigma$ upper-limit. Data from \\cite{Hartman1999:3eg}.} \\label{Fig:3c454:egret} \\end{figure} In 2005, \\source{} underwent a major flaring activity in almost all energy bands (see \\citealt{Giommi2006:3C454_Swift}). In the optical, it reached $R=12.0$\\,mag \\citep{vil06} and it was detected by \\igr{} at a flux\\footnote{Assuming a Crab-like spectrum.} level of $\\sim 3 \\times 10^{-2}$\\,photons\\,cm$^{-2}$\\,s$^{-1}$ in the 3--200~keV energy band \\citep{Pian2006:3C454_Integral}. Since the detection of the exceptional 2005 outburst, several monitoring campaigns were carried out to follow the source multifrequency behavior \\citep{vil06,vil07,rai07,rai08a,rai08b}. During the last of these campaigns, 3C~454.3 underwent a new optical brightening in mid July 2007, which triggered observations at all frequencies, including the \\agile{} one. In the following, we briefly describe the \\agile{} satellite, and we discuss the various campaigns triggered on \\source{}. ", "conclusions": "Since July 2007, \\source{} has been playing the same role for \\agile{} as 3C~279 had for EGRET, and during the period July 2007 - January 2009 we acquired data not only in the \\gray energy band, but across 14 decades in energy. This allowed us to construct simultaneous SEDs, sampling high, intermediate, and low \\gray emission states, involving both ground and space based observatories. We found that the role of the external Compton on the disk and the broad-line region radiation (and possibly also on hot corona photons) is crucial to account for the hard \\gray spectrum states. Moreover, thanks to the extremely dense optical coverage provided by the GASP-WEBT, we were able to study the correlations between the \\gray and optical fluxes. We found a $\\simeq 1$ day possible lag of the high energy photons with respect to the optical ones. The simultaneous presence of two \\gray satellites, \\agile{} and Fermi, the extremely prompt response of wide-band satellites as \\swi{}, and the long-term monitoring provided from the radio to the optical by the GASP-WEBT Consortium will assure the chance to investigate and study the physical properties of \\source{} and of several more blazars both at high and low emission states." }, "0910/0910.4323_arXiv.txt": { "abstract": "As of today over 40 planetary systems have been discovered in binary \\langeditorchanges{star systems}. In all cases the configuration appears to be circumstellar, where the planets orbit around one of the stars, the secondary acting as a perturber. The formation of planets in binary \\langeditorchanges{star systems} is more difficult than around single stars due to the gravitational action of the companion on the dynamics of the protoplanetary disk. In this contribution we first \\langeditorchanges{briefly} present the relevant observational evidence \\langeditorchanges{for} planets in binary systems. Then the dynamical influence that a secondary companion has on a circumstellar disk will \\langeditorchanges{be} analyzed through \\langeditorchanges{fully} hydrodynamical simulations. We demonstrate that the disk becomes eccentric and shows a coherent precession around the primary star. Finally, fully hydrodynamical simulations of evolving protoplanets embedded in disks in binary \\langeditorchanges{star systems} are presented. We investigate how the orbital evolution of protoplanetary embryos and their mass growth from cores to massive planets might be affected in this very dynamical environment. We \\langeditorchanges{consider, in particular,} the planet orbiting the primary in the system $\\gamma$ Cephei. ", "introduction": "Planet formation is obviously a process that occurs around single as well as in multiple \\langeditorchanges{star systems}, a fact that is indicated by the detection of well over 40 planetary systems that reside in a binary or even multiple star configurations. All of the observed systems display a so called S-type configuration in which the planets orbit around one of the stars and the additional star, the companion or secondary star, acts as a perturber to this system. In this review we shall refer to the secondaries as single objects, even though they may be multiple. As indicated in Table~\\ref{kley-tab:systems} the distances of the secondaries \\langeditorchanges{from} the host \\langeditorchanges{stars} of the planetary \\langeditorchanges{systems} range from very small values of about 20 AU for Gl~86 and $\\gamma$ Cep to several thousand AU. There are now 4 confirmed systems with a binary separation in the order of 20 AU; in \\langeditorchanges{Table~}\\ref{kley-tab:systems} \\langeditorchanges{these are} shown below the horizontal separation line. The mere existence of these 4 \\langeditorchanges{systems represents} a special challenge to any kind of planet formation process, due their tightness. Interestingly, there appears to be a lack of planets for \\langeditorchanges{intermediate separations} as there are no planets in binaries with separations between 20 and 100 AU. There are many more systems with larger \\langeditorchanges{separations} (not listed in the table), but in most of the cases only projected distances can be given, and the real physical \\langeditorchanges{separations are necessarily} larger. \\begin{table}[h,t] \\label{kley-tab:systems} \\begin{center} \\begin{tabular}{|l|l|l|l|l|l|} \\hline Star & $a_\\mathrm{bin}$ {\\small{[AU]}} & $a_\\mathrm{p}$ {\\small{[AU]}} & $M\\mathrmpl\\sin$ i {\\small{ [M$_{\\rm Jup}$]}} & $e_\\mathrm{p}$ & Remarks \\\\ \\hline HD 40979 & 6400 & 0.811 & ~3.32 & .23 & \\\\ Gl 777 A & 3000 & 3.65 & ~1.15 & .48 & \\\\ HD 80606 & 1200 & 0.439 & ~3.41 & .93 & \\\\ 55 Cnc B & 1065 & 0.1-5.9 & ~0.8-4.05 & .02-.34 & {\\small multiple} \\\\ 16 Cyg B & 850 & 1.66 & ~1.64 & .63 & \\\\ $\\upsilon$ And & 750 & 0.06-2.5 & 0.7-4.0 & .01-.27 & {\\small multiple} \\\\ HD 178911 B & 640 & 0.32 & ~6.3 & .12 & \\\\ HD 219542 B & 288 & 0.46 & ~0.30 & .32 & \\\\ $\\tau$ Boo & 240 & 0.05 & ~4.08 & .02 & \\\\ HD 195019 & 150 & 0.14 & ~3.51 & .03 & \\\\ HD 114762 & 130 & 0.35 & 11.03 & .34 & \\\\ HD 19994 & 100 & 1.54 & ~1.78 & .33 & \\\\ \\hline HD 41004A & 23 & 1.33 & ~2.5 & .39 & {\\small multiple} \\\\ $\\gamma$ Cep & 20.2 & 2.04 & 1.60 & .11 & {\\small $e_\\mathrm{bin}= 0.4$} \\\\ {HD 196885} & 17 & 2.63 & 2.96 & .46 & {\\small $e_\\mathrm{bin}= 0.4$} \\\\ Gl 86 & 20 & 0.11 & 4.0 & ~0.046 & {\\small White Dwarf} \\\\ \\hline \\end{tabular} \\end{center} \\caption{Some observed planets in binary \\langeditorchanges{star systems}. This is a selection with emphasis on the shorter period binaries (see \\, \\cite{2004A&A...417..353E,2006ApJ...646..523R,2008A&A...479..271C}). \\langeditorchanges{The} list is very incomplete for larger separations. } \\end{table} Despite the actual detection of planets in binary systems there is additional circumstantial evidence of debris disks (which are thought to be a byproduct of the planet formation process) in binary systems as indicated by Spitzer data. Here, for S-type configurations it is found that disks around an individual star of the binary exist mainly for binary separations larger than 50 AU, while P-type circumbinary debris disks are detected only in very tight binaries with $a_\\mathrm{bin}$ smaller than about 3 AU (\\cite{2007ApJ...658.1289T,2008ApJ...674.1086T,2008arXiv0808.1765Z}). As first pointed out by \\citet{2004A&A...417..353E}, {see also these proceedings}, there is statistical evidence for two interesting features in the mass-period and eccentricity period distribution of planets residing in binary systems: \\langeditorchanges{planets} with periods smaller than about 40 days tend to have larger masses than their counterparts in single star systems, while at the same time their eccentricities are smaller. This trend has been supported by \\langeditorchanges{the} more recent findings of \\citet{2007A&A...462..345D} who tried to correlate this with the tightness of the \\langeditorchanges{binary,} but the statistics \\langeditorchanges{are still based on small sample sizes} and more data are required. As the influence of the secondaries on the planet formation process will obviously be smaller for larger distances, we shall focus in this contribution on the more challenging tighter binaries and have used the physical \\langeditorchanges{parameters} of the $\\gamma$ Cep system for our models. Interestingly, $\\gamma$ Cep was one of the very first \\langeditorchanges{stars} which has been suggested to contain an extrasolar planet (of 1.7 $M_\\mathrm{Jup}$): {\\it ``This star has the firmest evidence of a very low mass companion''} \\citep{1988ApJ...331..902C}. A statement unfortunately retracted later by the same team \\citep{1992ApJ...396L..91W}, only to be rediscovered by \\citet{2003ApJ...599.1383H}. Today, this system is one of the tightest binary system known to contain a Jupiter-sized protoplanet. For this reason, it has attracted much attention in past years. Several studies looked at the stability and/or the possibility of (additional) habitable planets in the system \\citep{2004RMxAC..21..222D,2004MSAIS...5..127T,2006ApJ...644..543H,2006MNRAS.368.1599V}. In our studies we have taken the data for $\\gamma$~Cep from \\citet{2003ApJ...599.1383H}. The more recent data by \\citet{2007A&A...462..777N} \\langeditorchanges{only slightly} change the dynamical status. ", "conclusions": "In this contribution we have concentrated on the planetary growth process in relatively tight binary stars with particular \\langeditorchanges{attention given} to the system $\\gamma$~Cep. To study the effect of the binary we have followed the evolution of planetary embryos interacting with the ambient protoplanetary \\langeditorchanges{disk,} which is perturbed by the secondary star. As suspected, the perturbations of the disk, in particular its non-zero eccentricity and the periodic creation of strong tidally induced spiral density arms, lead to non-negligible \\langeditorchanges{effects} on the planetary orbital elements. While embryos placed in the disk at different initial distances from the primary star continue to migrate inwards at approximately the same rate, the eccentricity evolution is markedly different for the \\langeditorchanges{different} cases. If the initial distance is beyond about $a \\gsim 2.7$ AU the eccentricity of the embryo continues to rise to very high \\langeditorchanges{values,} and apparently \\langeditorchanges{the orbit remains bound only due to the damping action of the disk.} The main excitation mechanism of the initial rise of the eccentricity is the perturbed disk and the spiral arms near the outer edge of the disk. For a disk mass of 3$M_\\mathrm{Jup}$ a $1.6 M_\\mathrm{Jup}$ planet can easily be grown, and the final semi-major axis and eccentricity are also in the observed range of the $\\gamma$~Cep planet for suitable accretion rates onto the planet. One of the major problems in forming a planet in such a close binary system via the core instability model is the problem of the formation of the planetary core in the first place. Due to the large relative velocities induced in a planetesimal \\langeditorchanges{disk,} especially for objects of different \\langeditorchanges{sizes,} the growth process is also problematic in itself. Hence, the formation of the \\langeditorchanges{Jupiter-sized} planet observed in $\\gamma$~Cep via the standard scenarios remains difficult but may not be impossible. Future research will have to concentrate on additional physical effects such as radiative transport, three-dimensional effects and self-gravity of the disk." }, "0910/0910.3735_arXiv.txt": { "abstract": "This paper presents a detailed kinematic and chemical analysis of 16 members of the Kapteyn moving group. The group does not appear to be chemically homogenous. However, the kinematics and the chemical abundance patterns seen in 14 of the stars in this group are similar to those observed in the well-studied cluster, $\\omega$ Centauri. Some members of this moving group may be remnants of the tidal debris of $\\omega$ Cen, left in the Galactic disk during the merger event which deposited $\\omega$ Cen into the Milky Way. ", "introduction": "The idea of hierarchical galaxy formation, in which galaxies are believed to form from the aggregation of smaller elements (see review by \\citet{Freeman02}), has been around since \\citet{Searle78} first proposed this theory as a challenge to the belief that galaxies formed through the smooth collapse of a large protocloud \\citep{Eggen62}. The identification of debris from these smaller fragments remains of utmost importance in modern studies of theoretical and observational stellar dynamics. The most massive Galactic globular cluster, $\\omega$ Cen, has several unique physical properties which suggest that there are very significant differences in star formation histories, enrichment processes and structure formation between $\\omega$ Cen and other normal globular clusters \\citep{Bekki03}. A commonly accepted scenario of formation of $\\omega$ Cen is that it is the surviving nucleus of an ancient dwarf galaxy, the outer envelope of which was entirely removed by tidal stripping as it was accreted by the Galaxy (\\citet{Bekki03} and references therein). Through numerical simulations, \\citet{Bekki03} demonstrated the dynamical feasibility of $\\omega$ Cen forming from an ancient nucleated dwarf galaxy which was accreted into the young Galactic disk. \\citet{meza05} used numerical simulations to investigate the characteristics of tidal debris from satellite galaxies. They showed that these satellites deposit a large fraction of their stars into either the disc component of the Milky Way or into the halo, showing distinct ``trails\" in the angular momentum -- energy plane, depending on the plane of the satellite's orbit during disruption. Meza et al. discussed the presence of the $\\omega$ Cen stellar moving group (i.e. the stellar debris) in two studies of metal-poor stars in the solar neighbourhood, those of \\citet{Beers00} and \\citet{Gratton03}. Figure \\ref{mezafig} (taken from \\citet{meza05}) shows how the $\\omega$ Cen group is distinguished in the angular momentum distribution for the Gratton et al. (2003) sample. In both this sample and the Beers et al. (2000) sample, the $\\omega$ Cen group appears as an over-density of stars at very specific rotational velocity or angular momentum values. The radial (U) velocities for the $\\omega$ Cen candidate stars are observed to have a symmetric distribution, ranging from -300 to 300 km s$^{-1}$. An earlier simulation by \\citet{Bekki03} of the accretion of the $\\omega$ Cen parent galaxy showed a strong plume of debris stars in the solar neighborhood with L$_z$ near $-500$ kpc km s$^{-1}$. See also the discussion by \\citet{Mizutani03} on the kinematics of tidal debris from the parent galaxy. \\citet{dinescu02} also provided a theoretical prediction of where the $\\omega$ Cen group would appear kinematically using the metal-poor star sample of Beers et al. (2000). Figure \\ref{dinescu} shows the angular momentum -- energy plane, in which the shaded zone represents the area where expected $\\omega$ Cen candidate stars would lie. Dinescu defines this region by assessing three globular clusters ($\\omega$ Cen, NGC 362 and NGC 6779) that are believed to have come from the same original parent galaxy, and argues that their distribution may define the area in which further candidate remnants could lie. This region covers a large interval of energy over an angular momentum range from about L$_{z}$= -200 to -600 kpc km s$^{-1}$. We have recently investigated the chemical properties of stars in the Kapteyn moving group, first introduced by Eggen (1962). The stars of the Kapteyn group, as most recently tabulated by Eggen (1996), are mostly metal-poor and in retrograde galactic orbits, so they were identified as a halo moving group. Moving stellar groups can originate in several ways. Some form from a common gas cloud. As the resulting cluster disperses, its stars dissolve into the Galactic background yet maintain some common kinematical identity which may be used to identify members of a particular stellar group. Such moving groups represent a transition between bound clusters and field stars, and are probably chemically homogeneous (see \\citet{de-silva07} and references therein). Other moving groups appear to result from resonances in the galactic disk (eg \\citet{Dehnen98}), and others possibly as the debris of accreted and disrupted dwarf galaxies \\citep{Navarro04}. We were interested to see whether the Kapteyn group members were chemically similar, because this would be a pointer to the group's origin. It turned out that some of the group stars show chemical peculiarities similar to those seen in $\\omega$ Cen, and we will argue that the Kapteyn group may be part of the $\\omega$ Cen debris. Concentrations of metal-poor stars at L$_z$ values near those of $\\omega$ Cen and the Kapteyn group have appeared in several recent studies. \\citet{Dettbarn07} investigated substructure in a sample of stars with [Fe/H]$ < -1$ from the Beers et al (2000) catalog and identified a structure (their feature K) which appears to be related to the Kapteyn star group. The study of 246 metal-poor stars with accurate kinematics by \\citet{Morrison09} also showed a substantial number of stars at weakly retrograde values of L$_z$, similar to that of $\\omega$ Cen. \\citet{Klement09} find evidence of Kapteyn stream stars in their study of halo streams from the SDSS DR7. This paper presents a kinematic and chemical analysis of 17 members of the Kapteyn moving group. Section \\ref{kins} presents a kinematic analysis of the group. In Section \\ref{obs} the observations and reduction procedures are outlined, and Section \\ref{abunds} gives the details of the chemical abundance study. The summary and conclusions are given in Section \\ref{sumandconc}. ", "conclusions": "\\label{sumandconc} The kinematic and abundance analysis in this study suggests that at least our 14 members of the Kapteyn group and potentially many more stars in retrograde orbits which were not observed in this study, could be remnants of tidal debris stripped from the parent galaxy of $\\omega$ Cen, or even from the cluster itself, during its merger with the Galaxy. Our study provides the first detailed chemical evidence of field stars that appear to be both kinematically and chemically related to $\\omega$ Cen. It may lend weight to the view that $\\omega$ Cen is the remnant nucleus of a disrupted dwarf galaxy which was accreted by the Milky Way, by providing chemical evidence of tidal debris among the Galactic field stars. The three-banded structure seen in the Lindblad diagrams of Figures 3 and described in Section 2.1 may be indicative of stars shed with different energies from different wraps of the decaying orbit of the parent galaxy around the Milky Way. The reader is referred to Figure 6 in Meza et al. (2005) which shows several distinct E--L$_z$ curves from their numerical simulations of merger debris. In this study there is currently no evidence for an abundance difference between stars of different wraps of the orbit. A study involving higher quality data and a larger number of stars from these wraps is underway. The presence of this banded structure, along with the present Galactic radius of $\\omega$ Cen, suggest the original parent galaxy was relatively massive in order for dynamical friction to have the required effect, and was also relatively dense in order to survive the Galactic tidal stresses in to the current orbital radius of $\\omega$ Cen. What else can we infer about the parent galaxy? If our stars are debris from the parent galaxy, then the galactic metallicity-luminosity relationship and the mean metallicity of the sample ($\\langle {\\rm [Fe/H]} \\rangle = -1.5$) indicate that the parent galaxy's luminosity would have been about $M_V \\sim -11$. Its luminosity would have been about $2 \\times 10^6$ L$_\\odot$, and its stellar mass $\\sim 4 \\times 10^6$ M$_\\odot$. This stellar mass is comparable to the present stellar mass of $\\omega$ Cen itself. The total mass of the parent galaxy could have been as high as $5 \\times 10^8$ M$_\\odot$ if it were a dwarf galaxy with a dark matter content similar to the Fornax dSph (e.g \\citet{Walker06}). What are the chances of finding $\\omega$ Cen debris stars in the solar neighborhood from such a low-mass galaxy? We can use the Meza et al. simulation as a guide, although a detailed comparison is not appropriate as these authors note, because their end-product galaxy is not like the Milky Way. In summary, the debris of their satellite is well mixed azimuthally by redshift 0.48, at least within 10 kpc radius from the center of their parent galaxy, and is confined to a disk-like layer. The satellite is disrupted in its last three perigalactic passages, leaving a significant amount of substructure in E and L$_z$. We focus on the lowest $L_z$ substructure in their simulation, as a qualitative counterpart to the proposed $\\omega$ Cen debris. Most of our stars lie in a volume of radius $\\sim 500$ pc around the sun, and in a broad range of retrograde L$_z$. For comparison, we can estimate the fraction of the Meza et al. debris which would lie within our $500$ pc volume and within their lowest L$_z$ substructure: it is about $1.5 \\times 10^{-4}$. We could therefore expect to find about 700 debris stars in our volume and within the lowest angular momentum substructure, if we use the Meza et al. simulations as a guide and assume that the typical star in this population has a mass of about 0.8 M$_\\odot$. In the Dinescu region shown in Figure 3, we have compiled a total of about 100 stars. Even if all of them turn out to be $\\omega$ Cen debris stars, then this comparison may indicate that there are plenty more $\\omega$ Cen debris stars to be found nearby. We should also consider the possibility that the Kapteyn group stars were stripped from the cluster itself as its outer regions were disrupted by the Galactic tidal field. This would be the immediate inference from the the chemical similarity which we have demonstrated between the Kapteyn group stars and Cen stars. This seems unlikely, because the in-spiralling time under the influence of dynamical friction is in excess of the Hubble time for a cluster with the mass of $\\omega$ Cen; also, we note that the Kapteyn group stars are significantly more energetic than the cluster itself. It seems likely that the group stars came from the body of the parent dwarf galaxy, and therefore that the chemical peculiarities of $\\omega$ Cen are shared by its parent, and may well have originated in gas which was funneled into the cluster, possibly over an extended period, as suggested by \\citet{Norris97}. Detailed numerical models of the chemical and dynamical evolution of $\\omega$ Cen as its galaxy loses mass in the Galactic tidal field are needed. {\\bf Acknowlegements} ECW is funded by Australian Research Council grant DP0772283. KCF acknowledges his late colleagues Olin Eggen and Alex Rodgers for their work on the Kapteyn's star group which prompted this paper. We would also like to add our thanks to the referee (D. Casetti-Dinescu) for many helpful comments and suggestions.. We thank Mike Bessell for his major improvements to the echelle spectrograph which made this work possible; Daniela Carollo for discussions which led to the association of the Kapteyn's star group with $\\omega$ Cen and for comments on an earlier draft of this paper; John Norris and Tim Beers for discussions and encouragement; and the SSO staff for maintaining the SSO instrumentation." }, "0910/0910.3997_arXiv.txt": { "abstract": "Non-local thermodynamic equilibrium (NLTE) line formation for neutral and singly-ionized iron is considered through a range of stellar parameters characteristic of cool stars. A comprehensive model atom for Fe~I and Fe~II is presented. Our NLTE calculations support the earlier conclusions that the statistical equilibrium (SE) of Fe~I shows an underpopulation of Fe~I terms. However, the inclusion of the predicted high-excitation levels of Fe~I in our model atom leads to a substantial decrease in the departures from LTE. As a test and first application of the Fe~I/II model atom, iron abundances are determined for the Sun and four selected stars with well determined stellar parameters and high-quality observed spectra. Within the error bars, lines of Fe~I and Fe~II give consistent abundances for the Sun and two metal-poor stars when inelastic collisions with hydrogen atoms are taken into account in the SE calculations. For the close-to-solar metallicity stars Procyon and $\\beta$~Vir, the difference (Fe~II - Fe~I) is about 0.1~dex independent of the line formation model, either NLTE or LTE. We evaluate the influence of departures from LTE on Fe abundance and surface gravity determination for cool stars. ", "introduction": "Iron plays an outstanding role in studies of cool stars thanks to quite numerous lines in the visible spectrum, which are easy to detect even in very metal-poor stars. Iron serves as a reference element for all astronomical research related to stellar nucleosynthesis and chemical evolution of the Galaxy. Iron lines are used to determine the surface gravity, $\\logg$, and the microturbulence $\\xi$ of stellar atmospheres. In the atmosphere with $\\Teff > 4500$~K, neutral iron is a minority species. The ionization equilibrium between Fe~I and Fe~II and the excitation equilibrium of Fe~I easily deviate from thermodynamic equilibrium. Since the beginning of the 1970s a number of studies attacked the problem of non-local thermodynamic equilibrium (NLTE) for Fe (e.g., \\cite[Athay \\& Lites (1972)]{Athay1972}, \\cite[Thevenin \\& Idiart (1999)]{Thevenin1999}, \\cite[Gehren \\etal\\ (2001)]{Gehren2001}). However, a consensus on the expected magnitude of the NLTE effects was not reached. In this study, we update the model atom of Fe I-II treated by \\cite[Gehren \\etal\\ (2001)]{Gehren2001} (hereafter Paper~I) and apply it to analysis of the Fe spectrum in the Sun and selected cool stars with the aim of empirically constraining the role of inelastic collisions with hydrogen atoms in the SE of Fe~I-II. \\firstsection ", "conclusions": "\\begin{itemize} \\item Completeness of model atom for Fe~I is important for a correct calculation of the Fe~I/Fe~II ionization equilibrium in the atmosphere of cool stars. \\item Thermalizing processes not involving electron collisions have to be included in the SE calculations for Fe~I-II. Collisions with hydrogen atoms could be good candidates for such processes. \\item Fe I is affected by significant NLTE effects for giants and very metal-poor stars. \\item Only minor departures from LTE are obtained for Fe II. \\end{itemize} \\bigskip {\\it Acknowledgements.} L.M. acknowledges a partial support from the International Astronomical Union, the Russian Foundation for Basic Research (08-02-92203-GFEN), and the Russian Federal Agency on Science and Innovation (02.740.11.0247) of the participation at the IAU XXVII General Assembly. This study is supported by the Deutsche Forschungsgemeinschaft (GE 490/34.1). A.K. acknowledges support by the Swedish Research Council (VR)." }, "0910/0910.3445_arXiv.txt": { "abstract": "We present local extrema studies of two models that introduce a preferred direction into the observed cosmic microwave background (CMB) temperature field. In particular, we make a frequentist comparison of the one- and two-point statistics for the dipole modulation and ACW models with data from the five-year \\textit{Wilkinson Microwave Anisotropy Probe} (\\WMAP). This analysis is motivated by previously revealed anomalies in the \\WMAP\\ data, and particularly the difference in the statistical nature of the temperature anisotropies when analysed in hemispherical partitions. The analysis of the one-point statistics indicates that the previously determined hemispherical variance difficulties can be apparently overcome by a dipole modulation field, but new inconsistencies arise if the mean and the $\\ell$-dependence of the statistics are considered. The two-point correlation functions of the local extrema, $\\xi_{\\rm TT}$ (the temperature pair product) and $\\xi_{\\rm PP}$ (point-point spatial pair-count), demonstrate that the impact of such a modulation is to over-`asymmetrise' the temperature field on smaller scales than the wave-length of the dipole or quadrupole, and this is disfavored by the observed data. The results from the ACW model predictions, however, are consistent with the standard isotropic hypothesis. The two-point analysis confirms that the impact of this type of violation of isotropy on the temperature extrema statistics is relatively weak. From this work, we conclude that a model with more spatial structure than the dipole modulated or rotational-invariance breaking models are required to fully explain the observed large-scale anomalies in the \\WMAP\\ data. ", "introduction": "\\label{sec_intro} Local extrema in the CMB have been extensively studied in the context of hotspots (peaks) and coldspots (troughs) arising in Gaussian random fields \\citep{bond_etal_1987, vittorio_etal_1987}. Additional statistics related to such local extrema, such as the Gaussian curvature and temperature-correlation function, have been investigated as a means to distinguish the geometry of the universe and test the Gaussian hypothesis for the nature of the initial conditions \\citep{barreiro_etal_1997, heavens_etal_1999, heavens_etal_2001}. Such statistical techniques have subsequently been applied to several datasets, and most notably by \\citet{kogut_etal_1995, kogut_etal_1996} with the $COBE$-DMR data. Observations from the \\textit{Wilkinson Microwave Anisotropy Probe} (\\WMAP) currently provide the most comprehensive, full-sky, high-resolution CMB measurements to-date, and studies on the local extrema properties thereof have been undertaken \\citep{LW04, LW05, tojeiro_etal_2006}. More recently in \\citet{hou_etal_2009}, we extensively analyzed the statistical properties of both the one- and two-point statistics of local extrema in the five-year \\WMAP\\ temperature data. Such extrema are defined as those pixels whose temperature values are higher (maxima) or lower (minima) than all of the adjacent pixels \\citep{wandelt_etal_1998}. We considered only that part of the sky outside of the \\WMAP\\ KQ75 mask and its subsequent partition in either Galactic or Ecliptic coordinates into northern and southern hemispheres (hereafter GN, GS, EN, ES). A frequentist comparison with the predictions of a Gaussian isotropic cosmological model that adopted the best-fit parameters from the \\WMAP\\ team was then made. The hypothesis test indicated a low-variance of both local maxima and minima in the Q-, V- and W-band data that was inconsistent with the Gaussian hypothesis at the 95\\% C.L. The two-point analysis showed that the observed temperature pair product at a given threshold $\\nu$, $\\xi_{\\rm TT}(\\theta,\\nu > 1,2)$, indicates a $3\\sigma$ level `suppression' on GN and EN, whereas $\\xi_{\\rm TT}(\\theta<20^{\\circ},\\nu < \\infty)$ is suppressed on the full-sky and both northern hemisphere partitions. The latter is also the case for the point-point spatial pair-count function $\\xi_{\\rm PP}(\\theta<20^{\\circ},\\nu > 1,2)$. Intriguingly, the statistics showed an $\\ell$-dependence such that consistency with the Gaussian hypothesis was achieved once the first 5 or 10 best-fitting multipoles were subtracted, implying that the anomalies may be connected to features of the large-scale multipoles. The local extrema anomalies therefore provide further evidence of a hemispherical asymmetry that was originally revealed using the power spectrum \\citep{hansen_etal_2004, eriksen_etal_2004, hansen_etal_2008}, the N-point correlation functions \\citep{eriksen_etal_2005} and Minkowski functionals \\citep{park_2004}. This can be interpreted as a violation of the cosmological principle of isotropy. Theoretically, \\citet{gordon_etal_2005} proposed a mechanism of spontaneous isotropy breaking in which the long-wavelength modes of a mediating field couple non-linearly to the CMB perturbations. These fluctuations appear locally as a gradient and imprint a preferred direction on the sky. An implementation of a multiplicative modulation field of the intrinsic anisotropy with a single preferred direction was fitted to the three-year \\WMAP\\ observations in \\citet{eriksen_etal_2007} . The so-called \\lq dipole modulation field' (dmf) was detected at a significant confidence level, and confirmed at even higher significance in the five-year \\WMAP\\ data by \\citet{hoftuft_etal_2009}. Another mechanism that violates rotational invariance was proposed by \\citet{acw_2007}. The ACW model picks out a preferred direction during cosmological inflation to modify the power spectrum of primordial perturbations, $P(k)$. Its imprint on the CMB can then be described by the covariance matrix of spherical harmonic coefficients, $\\langle a_{\\ell m}a^{*}_{\\ell'm'}\\rangle$. \\citet{groe_hke_2009} estimated the parameters of the ACW model in an extended CMB Gibbs sampling framework from the five-year \\WMAP\\ data. The posterior distribution of the parameters indicated a convincing detection of isotropy violation at $3.8\\sigma$ significance in the W-band data. In this paper, we make a frequentist comparison of the dmf and ACW models with the five-year \\WMAP\\ maps of temperature anisotropy using a large number of Monte-Carlo simulations. In particular, we consider the one- and two-point statistics of local extrema and evaluate whether the \\WMAP\\ data are more consistent with these models as opposed to the standard Gaussian cosmological models against which various anomalies have been claimed. A rigorous hypothesis test methodology is applied to establish the significance at which the models impact the previous conclusions regarding \\WMAP\\ data and the violation of isotropy. In addition, the $\\xi_{\\rm TT}$ and $\\xi_{\\rm PP}$ correlation functions are utilised to further confirm the findings of the one-point analysis and to study the scale-dependence of the local extrema both in real and spherical harmonic space. It might be considered that the statistical significance of the local extrema anomalies alone is insufficient to warrant an analysis of the anisotropic models, in particular given the additional parameters required to specify them. However, that they provide significantly better fits to the \\WMAP\\ data than the isotropic model has been demonstrated by independent Bayesian power spectrum analysis. It is clearly then quite legitimate to consider the local extrema statistics in terms of these improved models, and to determine whether the observed local extrema are also more consistent with such cosmological prescriptions. In fact, we will find that there is little evidence that the extrema anomalies are remedied by these models, and indeed in one case we show that the anomalous behaviour becomes more significant. This paper is organised as follows. In Section~\\ref{subsec_wmap_data}, we present an overview of the \\WMAP\\ data and the instrumental properties that must be involved in Gaussian simulations to enable an unbiased comparison with the real data. The two models and the algorithms for simulating them are briefly reviewed in Section~\\ref{subsec_2model}. Section~\\ref{subsec_analysis} prescribes the technique used for further data-processing and introduces the one- and two-point analysis methods. The results are reported in Section~\\ref{sec_results} before we present our conclusions in Section~\\ref{sec_conclusion}. ", "conclusions": "\\label{sec_conclusion} In this paper, the statistical properties of local temperature extrema in the 5-year \\WMAP\\ data have been compared with two models that break rotational invariance -- dipole modulation and the ACW model. Such a comparison was motivated by the fact that our previous analysis of local extrema statistics \\citep{hou_etal_2009} demonstrated an unlikely hemispherical asymmetry in the GN and EN as a consequence of an extremely low-variance compared to the expectations of a Gaussian isotropic scenario. Moreover, the 5-year \\WMAP\\ data indicate significant detections of these models on the basis of power spectrum analyses. We employ Gaussian MC simulations encoding the features of these two models to establish the statistical basis for testing whether the breaking of rotational invariance by the models can afford a satisfactory explanation of local extrema anomalies. As in previous work, both one and two-point statistics, as well as their dependence on large angular-scale modes have been studied. Both models are parameterised by a set of amplitudes and preferred directions as established by independent Bayesian analyses in \\citet{hoftuft_etal_2009} and \\citet{groe_hke_2009}. Our analysis has been carried out by sampling the former within their 95\\% confidence intervals, whilst imposing the fixed preferred direction found for each band, since the latter are estimated to be robust for each model. The V and W-band data are considered here. In fact, the local extrema studies based on one and two-point statistics indicate that neither model provides a satisfactory solution to observed properties. In particular, the ACW model is probably the least interesting in this context -- both the one-point analysis and two-point studies show similar features to the isotropic results, including their $\\ell$-dependence. The dipolar modulation field, however, may itself be constrained by our analysis. Results determined from division of the data into hemispheres implies that a dmf with significant amplitude may suppress the observed variance anomalies. In particular, the $\\lrmv=2$ values on EN and ES hemispheres indicates consistency of the data with the model. However, after subtraction of the first 10 multipoles, new problems arise in the southern hemispheres, where the observed variance of local extrema is now much lower than the model prediction. This is suggestive that the dmf may not have the correct profile over the full-sky in order to reconcile the variance properties of the observed local extrema with the Gaussian anisotropic simulations, and that the amplitude of the effect may also be a function of $\\ell$. Moreover, the observed mean statistics, which showed few anomalies when compared to the Gaussian isotropic model, are not consistent and contradict the variance results by requiring a weaker amplitude. Our analysis indicates that neither a simple dipole modulated field model nor a rotational-invariance breaking model can satisfactorily explain the local extrema anomalies present in the \\WMAP\\ data. On the one side, the rotational-invariance model affects the large-scale temperature amplitudes too little to significantly affect the local extrema statistics. On the other side, a modulation type model needs a more elaborate spatial structure than a simple dipole to fully fit the data. These issues remain interesting for further investigation." }, "0910/0910.5169_arXiv.txt": { "abstract": "We perform three-dimensional relativistic hydrodynamical simulations of the coalescence of strange stars and explore the possibility to decide on the strange matter hypothesis by means of gravitational-wave measurements. Self-binding of strange quark matter and the generally more compact stars yield features that clearly distinguish strange star from neutron star mergers, e.g. hampering tidal disruption during the plunge of quark stars. Furthermore, instead of forming dilute halo structures around the remnant as in the case of neutron star mergers, the coalescence of strange stars results in a differentially rotating hypermassive object with a sharp surface layer surrounded by a geometrically thin, clumpy high-density strange quark matter disk. We also investigate the importance of including nonzero temperature equations of state in neutron star and strange star merger simulations. In both cases we find a crucial sensitivity of the dynamics and outcome of the coalescence to thermal effects, e.g. the outer remnant structure and the delay time of the dense remnant core to black hole collapse depend on the inclusion of nonzero temperature effects. For comparing and classifying the gravitational-wave signals, we use a number of characteristic quantities like the maximum frequency during inspiral or the dominant frequency of oscillations of the postmerger remnant. In general, these frequencies are higher for strange star mergers. Only for particular choices of the equation of state the frequencies of neutron star and strange star mergers are similar. In such cases additional features of the gravitational-wave luminosity spectrum like the ratio of energy emitted during the inspiral phase to the energy radiated away in the postmerger stage may help to discriminate coalescence events of the different types. If such characteristic quantities could be extracted from gravitational-wave signals, for instance with the upcoming gravitational-wave detectors, a decision on the strange matter hypothesis and the existence of strange stars should be possible. ", "introduction": "The equation of state (EoS) of high-density matter and therefore the true nature of compact stars has been a mystery since the discovery of these objects. Besides the possibility of a neutron-proton composition more exotic phases have been proposed to appear in the cores of compact stars (see e.g. \\cite{2007ASSL..326.....H} for a review, for early considerations of quark stars see~\\cite{1969NCimL...2...13I,1970PThPh..44..291I}). The formulation of the strange matter hypothesis \\cite{PhysRevD.4.1601,PhysRevD.30.272} introduced the possibility that compact stars are so-called strange stars (SSs) \\cite{1986A&A...160..121H,1986ApJ...310..261A}. According to this hypothesis the absolute ground state of matter is formed by a quark phase consisting of up, down and strange quarks, and compact stars are self-bound objects composed of this kind of matter. It turns out that SSs are in many ways similar to neutron stars (NSs), e.g. in the mass range and compactness, and therefore they can be considered as an alternative explanation for compact stars \\cite{1986A&A...160..121H,1986ApJ...310..261A,1996csnp.book.....G,2007ASSL..326.....H,2005PrPNP..54..193W}. Throughout this article the term compact star refers to either a NS or a SS. Theoretically as well as observationally, the question whether the strange matter hypothesis is true and whether (at least some) NSs are actually SSs, is an open issue. Besides the determination of compact star parameters, several possibilities have been proposed to explore the observational consequences of the strange matter hypothesis, including the possibility that small lumps of this strange quark matter (SQM) with baryon numbers of about $10^2$ and more might be abundant in the cosmic ray flux (see e.g. \\cite{2005PhRvD..71a4026M}). Experiments like AMS-02, which is planned to be installed on the International Space Station in 2010, have the potential to detect these so-called strangelets \\cite{ams,2004JPhG...30S..51S}. Also, SQM might be detected with the LHC at CERN \\cite{SchaffnerBielich:1996eh,Greiner:1987tg,Spieles:1996is}. For a review on these considerations and other observational implications of the strange matter hypothesis see \\cite{2005PrPNP..54..193W,2006JPhG...32S.251F}. With the current and upcoming GW detectors like LIGO \\cite{:2007kva} and VIRGO \\cite{Acernese:2006bj} and the prospect of a detection of signals from compact object binaries \\cite{2004ApJ...601L.179K}, the question naturally arises if the signal of mergering NSs could be distinguished from SS mergers, especially because compact object binaries are considered to be among the most promising sources for these detectors. The systematic investigation of the imprint of the EoS on the gravitational-wave (GW) signal is still in its infancy. Most studies consider simplified EoSs like polytropes and try to explore the chances to measure some general compact star properties like the stellar radius (e.g. \\cite{1996PhRvD..54.7261Z,2009arXiv0901.3258R}). It is not clear whether a decision on the strange matter hypothesis could be made on the basis of such measurements, because the compactness of SSs can be in the same range as that of NSs. First attempts of exploring the consequences of microphysical EoSs including nonzero temperature effects in a relativistic treatment, necessary for reliable results, have been made by \\cite{2007PhRvL..99l1102O}. Also fully relativistic studies have been conducted in which temperature corrections were approximated by an ideal gas component added to a zero-temperature microphysical EoS (e.g. \\cite{2006PhRvD..73f4027S} and preceding works). Strange stars as sources of GWs have been considered in a rotational equilibrium approach investigating the (final phase) of the inspiral of SS binaries \\cite{PhysRevD.71.064012,GondekRosinska:2008nf}. Quasiequilibrium orbits were constructed to determine the innermost stable circular orbit (ISCO). It was found that the frequency of the GWs at the ISCO depends on the compactness of the SSs. The orbital evolution and the associated GW emission of SS-black hole binaries has been estimated within a semianalytic model in \\cite{2004JPhG...30.1279P}. Moreover, the signals from various instabilities of rotating SSs have been worked out in \\cite{2000PhRvL..85...10M,2002MNRAS.337.1224A,GondekRosinska:2003iy} (see also references therein). In \\cite{2008PhRvL.101r1102F} it was found that the frequencies of the g-mode in newly born SSs are significantly lower than those of NSs. Differences of the fundamental pressure modes are discussed in \\cite{2007GReGr..39.1323B} where a discrimination was found only to be possible if additional information like the mass was available. Nevertheless it is not clear whether these discriminating features are relevant as the corresponding GW signals might be too weak for measurements. In this article we examine the GW signals emitted by merging SSs. We explore two different EoSs representing different possibilities of SQM properties and consider the inspiral phase, the final plunge, and the postmerger stage. We compare the GW characteristics of SS coalescence with those of ordinary NSs and discuss if GW observations could be decisive on the strange matter hypothesis. For that we conduct three-dimensional relativistic hydrodynamical simulations that allow us to follow the evolution of the whole merging process and to extract observational signatures. To our knowledge this is the first study considering these events in dynamical simulations; only results for mergers of SS-black hole systems have been so far reported in \\cite{2002MNRAS.335L..29K}, where Newtonian hydrodynamics were used and the black hole (BH) was modeled by a pseudorelativistic potential \\cite{1980A&A....88...23P}. The paper is structured as follows: Sect.~\\ref{sec:num} will introduce the underlying model and numerical methods. In Sect.~\\ref{sec:eos} the properties of the EoSs used in this study will be described. The simulations and some general aspects will be discussed in Sect.~\\ref{sec:sim}. In Sect.~\\ref{sec:GW} the GW signals will be presented and in Sect.~\\ref{sec:con} we will close with a summary and our conclusions. ", "conclusions": "\\label{sec:con} We have reported on simulations of SS mergers and discussed our results in comparison to NS coalescence events. In particular, we focussed on the GW signals and their potential to provide information about the existence of SSs. We found that the dynamical behavior of SS mergers is fundamentally different, which can be understood by the higher compactness of SSs and their merger remnants, the self-binding of SQM and the different influence of thermal effects. While NS mergers form a dilute halo or toruslike structure around a dense, hypermassive, differentially rotating remnant, the remnant of merging SSs is bounded by a sharp surface as were the initial stars. Only by the formation of thin tidal arms relatively late in the evolution of the remnant, matter gets shed off the central object and forms a fragmented thin disk in the equatorial plane. In order to estimate the importance of thermal effects during NS and SS mergers we compared our models with simulations where zero temperature was imposed by extracting the thermal energy as most efficient cooling would do. For both kinds of stars we observed a sensitive dependence of the dynamics of the system on thermal effects, affecting for instance the development and structure of the outer remnant parts of NS mergers. While the estimated relic torus mass after collapse of the remnant to a BH is in general higher if thermal effects were neglected, a nonzero temperature of NS matter leads to an inflated, dilute halolike torus in contrast to a much thinner, more disklike structure for $T=0$. For SS mergers we found a shorter time delay for the BH formation, whereas in the case of NSs this time interval is stretched as reported in \\cite{2008PhRvD..78h4033B}. This difference can be understood as a consequence of the additional gravitating effect of the thermal energy, which in the case of very dense SSs is not overcompensated by thermal pressure effects (different from the NS case). The analysis of the GW signals emitted by SS mergers revealed that already by means of relatively simple characteristic features of the signal it may be possible to decide whether SS or NS mergers produced the emission. In particular, we found that the maximal frequency during the inspiral and the frequency of the ringdown of the postmerging remnant are in general higher in the case of SS mergers in comparison to NS mergers. Whether this criterion can be used to finally decide on the strange matter hypothesis depends on the particular stellar properties associated with the EoS of high-density matter. For a similar mass-radius relation within a certain mass range, meaning relatively compact NSs or less compact SSs (the LS-MIT60 scenario), the determination of these frequencies might not be decisive. However, taking additional characteristics of the GW luminosity into consideration will allow for a discrimination. In this context we discussed the occurrence of a prompt collapse, the ratios of the GW energy emitted in the postmerger phase to the energy radiated away during the inspiral phase, the growth rate of the energy emission associated with the postmerger ringdown signal, and the appearance of a gap in the GW luminosity spectrum. In addition, it was shown in \\cite{2008arXiv0812.4248B} that cosmic ray experiments will yield information about the strange matter hypothesis and may clarify the situation, especially in the case when GW signals are not conclusive as in the LS-MIT60 scenario. Therefore we expect that despite the rather uncertain event rate the upcoming advanced GW detectors LIGO and VIRGO may provide valuable data to decide about this long-standing question, and it appears likely that one will thus gain fundamental information about the properties of high-density matter. Moreover, future GW detectors like the Einstein telescope \\cite{et} and the DUAL detector \\cite{dual} will have a higher sensitivity in the high-frequency domain above 1~kHz, where characteristic features occur. Signals measured with these future instruments will therefore allow for a more detailed analysis. Finally, we would like to point out that also other forms of self-bound matter have been discussed in the literature (see e.g.~\\cite{1977JETPL..25..465V,1990PhR...192..179M,2007ASSL..326.....H}). For instance pion condensates may lead to stellar objects similar to SSs, and also abnormal nuclei similar to strangelets may exist in this case. Therefore, the study presented in this work should be generalized to these states of matter, where quantitative considerations require the knowledge of the specific EoS. However, for stellar properties comparable to those of SSs we expect qualitatively the same behavior and the same GW features as for SSs. From this discussion it is clear that only a multimessenger approach can bring a final answer. In addition to GW measurements and cosmic ray experiments, this may in particular include the observation of isolated compact stars to derive the cooling history of these objects, which is significantly different for NSs, SSs and self-bound pion-condensed stars (see~\\cite{1999paln.conf.....W,2000ApJ...533..406B,2004ARA&A..42..169Y,2005PhRvC..71d5801G}). Several improvements and extensions may supplement this first study of SS mergers in the future. Besides refinements in the methodical approach like a fully relativistic treatment and the inclusion of magnetic fields (only important during the merging phase if very strong, but must be expected to grow afterwards and affect the ringdown phase) and radiative effects, one would also like to account for the effects of quark interaction and color superconductivity, employ models beyond the MIT bag model, and discuss other forms of self-bound matter. Moreover, a determination of the modes excited in the merger remnants may be a promising way to develop a better understanding of the origin of the characteristic properties of the GW spectra. Furthermore a study of SS-BH mergers will yield further insights about the behavior of such systems in contrast to NS-BH binaries." }, "0910/0910.2244.txt": { "abstract": "We discovered a sample of 250 Ly$\\alpha$ emitting (LAE) galaxies at $z\\simeq2.1$ in an ultra-deep 3727{\\AA} narrow-band MUSYC image of the Extended Chandra Deep Field-South. LAEs were selected to have rest-frame equivalent widths (EW) $>20$~{\\AA} and emission line fluxes $F_{Ly\\alpha} >2.0 \\times 10^{-17}$~erg~ cm$^{-2}$~s$^{-1}$, after carefully subtracting the continuum contributions from narrow-band photometry. The median emission line flux of our sample is $F_{Ly\\alpha} = 4.2 \\times10^{-17}$~erg~cm$^{-2}$~s$^{-1}$, corresponding to a median Ly$\\alpha$ luminosity L$_{Ly\\alpha} = 1.3 \\times10^{42}$~erg~s$^{-1}$ at $z\\simeq2.1$. At this flux our sample is $\\geq90$~\\% complete. Approximately 4\\% of the original NB-selected candidates were detected in X-rays by Chandra, and 7\\% were detected in the rest-frame far-UV by GALEX; these objects were eliminated to minimize contamination by AGN and low-redshift galaxies. At L$_{Ly\\alpha} \\geq 1.3 \\times10^{42}$~erg~s$^{-1}$, the equivalent width distribution is unbiased and is represented by an exponential with scale-length 83~$\\pm10$~{\\AA}. Above this same luminosity threshold, we find a number density of $1.5\\pm0.5\\times 10^{-3}$~Mpc$^{-3}$. Neither the number density of LAEs nor the scale-length of their EW distribution show significant evolution from $z\\simeq3$ to $z\\simeq2$. We used the rest-frame UV luminosity to estimate a median star formation rate of 4 M$_{\\odot}$~yr$^{-1}$. The median rest-frame UV slope, parametrized by the color $B-R$, is that typical of dust-free, 0.5-1 Gyr old or moderately dusty, 300-500 Myr old population. Approximately 30\\% of our sample is consistent with being very young (age~$<100$ Myr) galaxies without dust. Approximately 40\\% of the sample occupies the $z\\sim2$ star-forming galaxy locus in the $UVR$ two color diagram, but the true percentage could be significantly higher taking into account photometric errors. Clustering analysis reveals that LAEs at $z\\simeq2.1$ have $r_0=4.8\\pm0.9$ Mpc, corresponding to a bias factor $b=1.8\\pm0.3$. This implies that $z\\simeq2.1$ LAEs reside in dark matter halos with median masses log(M/M$_{\\odot})= 11.5^{+0.4}_{-0.5}$, which are among of the lowest-mass halos yet probed at this redshift. We used the Sheth \\& Tormen conditional mass function to study the descendants of these LAEs and found that their typical present-day descendants are local galaxies with L$^*$ properties, like the Milky Way. ", "introduction": "The search for high-redshift star-forming galaxies advanced rapidly with the introduction of the Lyman Break Galaxies (LBG) technique (\\citealt{Gu:1990}, \\citealt{StHm:1992}, \\citealt{Steidel:1999}) that takes advantage of the lack of flux at wavelengths shorter than the Lyman break at 912~{\\AA} ~rest frame due to absorption of ionizing photons by neutral hydrogen, located in stellar atmospheres, in the interstellar medium (ISM), and in the intergalactic medium (IGM) between galaxies. At $z=3$ the break is located in the observed U band and at higher redshift it moves into optical and infrared bands. This has allowed an exploration of star-forming galaxies at redshifts $3\\leq z\\leq8$ via imaging from ground and space (e.g. \\citealt{Steidel:2003}, \\citealt{Bouwens:2006}, \\citealt{Ouchi:2008}) and spectroscopy on 8--10 meter telescopes (e.g. \\citealt{Shapley:2001}, \\citealt{Shapley:2003}). A significant fraction of high redshift LBGs show the Ly$\\alpha$ line in emission (\\citealt{Shapley:2001}). This emission offers additional information about the process of star formation inside these galaxies and radiative transfer in their ISM. Looking for galaxies with Ly$\\alpha$ in emission has become an important photometric technique that permits us to find faint (R $\\sim$ 27) star-forming galaxies at high redshift. This technique consists of comparing the flux density measured in a narrow-band filter, revealing observed-frame Ly$\\alpha$ emission to that found in the broad-band filters, representing the continuum. Thanks to the intensity of this emission line, the resulting Ly$\\alpha$ emitting (LAE) galaxies provide a special population of high redshift galaxies. The properties of LAEs have been extensively studied at $z\\ge3$ (e.g. \\citealt{Ouchi:2005}, Venemans et al. 2005, \\citealt{Gawiser:2006b}, \\citealt{Gronwall:2007}, ~\\citealt{Nilsson:2007}). LAE samples are composed primarily of galaxies fainter in the continuum than LBGs; Ly$\\alpha$ Emitting galaxies therefore probe the lowest bolometric luminosities at high redshift. Theoretical models, that include radiative transfer inside star-forming galaxies (\\citealt{V:2006}, \\citealt{SV:2008}, \\citealt{V:2008}, \\citealt{Atek:2009}), were also developed to understand how Ly$\\alpha$ photons form in HII regions and then escape the galaxy, depending on resonant scattering by neutral hydrogen, dust absorption and velocity dispersion in the interstellar medium. The amount of dust and the interstellar medium geometry can affect the escape of Ly$\\alpha$ photons and hence the shape of the line. Clumpy media could permit Ly$\\alpha$ photons to escape, even if the galaxy is not dust-free (\\citealt{Neufeld:1991}, \\citealt{Fin:2008}, \\citealt{Fyougold:2009}). Spectral Energy Distribution (SED) fitting of the stacked multi-wavelength photometry of $z\\simeq3$ LAEs (\\citealt{Gawiser:2007}, \\citealt{Lai:2008}) shows they are a young (median starburst age of $\\sim20$ Myr), low stellar mass (M$\\sim10^9$ M$_{\\odot}$), modest SFR (median SFR $\\sim2$ M$_{\\odot}$~yr$^{-1}$), low dust (A$_V\\leq$0.2) population of galaxies in an active phase of star formation. SEDs have also shown older population best fits for subsamples of LAEs at $z>3$ (\\citealt{Pirzkal:2007}, \\citealt{Ono2009}, \\citealt{Nilsson:2009}). SED fitting of individual galaxies showed older ages, higher stellar mass and more dust for continuum-bright LAEs drawn from LBG samples (\\citealt{Shapley:2001}, \\citealt{Tapken:2007}, \\citealt{Pentericci:2009}). \\citet{Stiavelli2001} had also shown redder colors for some LAEs at $z\\simeq2.4$. Recently \\citealt{Nilsson:2009} presented the first results of observations of LAEs at $z\\simeq2.3$, inferring evolution in the properties from $z\\sim3$ to $z\\sim2$, with more diversity in photometric properties at $z\\simeq2.3$. Clustering analysis of LAEs showed $z\\ge4$ LAE bias factors (\\citealt{kovac:2007}, \\citealt{Ouchi:2004}) expected for progenitors of massive elliptical galaxies in the local Universe, while $z\\simeq3.1$ LAEs could be progenitors of L$^*$ galaxies (\\citealt{Gawiser:2007}). Semi-analytical simulations were also able to reproduce these results (\\citealt{Orsi2008}). Lower redshift observations, including clustering, will reveal evolution from high to low redshift. For this reason we were motivated to study LAE samples at redshift around 2. This will trace the star formation properties of this type of galaxy at the epoch of the peak of cosmic star formation density (\\citealt{Madau:1998}, \\citealt{Giav2004}). It also promises to reveal $z\\sim0$ descendants of LAEs at $z\\simeq2.1$. In this paper we describe the results from ultra-deep 3727~{\\AA} narrow-band MUSYC (MUlti-walengthSurvey Yale Chile, \\citealt{Gawiser:2006a}) imaging of the 998 arcmin$^2$ Extended Chandra Deep Field-South. In sections 2 and 3 we present the observations and the data reduction. In Section 4 we summarize the selection of the LAE sample and estimate the possible contaminants. We present the properties of the LAE sample in Section 5: number density, star formation rate, colors and clustering. In Section 6 we discuss the results and derive conclusions. We assume a $\\Lambda$CDM cosmology consistent with WMAP 5-year results (\\citealt{Dunkley:2009}, their table 2), adopting the mean parameters $\\Omega_m=0.26$, $\\Omega_{\\Lambda}=0.74$, H$_0=70$ km~sec$^{-1}$~Mpc$^{-1}$, $\\sigma_8=0.8$. ", "conclusions": "\\label{sec:discussion} We imaged the ECDF-S using a NB3727 narrow-band filter, corresponding to the wavelength of Ly$\\alpha$ emission at $z\\simeq2.1$. Following the formalism described in the Appendix, we applied the color cut $UB$$-$NB3727$ > 0.73$ and additional significance criteria that yielded a sample of 250 LAEs. In our observation we achieve the same 5$\\sigma$ detection limit in Ly$\\alpha$ luminosity (log(L(Ly$\\alpha$))=41.8) as the sample of LAEs at $z\\simeq3.1$ (Gronwall et al. 2007, Gawiser et al. 2007). Therefore we are able to look for indications of evolution between $z\\sim$2 - 3. Concentrating on $z\\sim2$, we compare LAEs with star-forming galaxies (Steidel's BX sample), which can also show the Ly$\\alpha$ line in emission. In many cases our analysis concentrates on the typical properties of the LAE sample as a whole; it is important to remember that there will always be cases of individual LAEs whose physical properties differ considerably from those of the typical LAE. The magnitude distribution of LAEs at $z\\simeq2.1$ (Fig.~\\ref{fig:mag}) is consistent with that predicted by the $z\\simeq3.1$ LAE Ly$\\alpha$ luminosity function, but with about twice the normalization, i.e. total number density. As reported in \\S 5.1 we calculated a LAE number density of $3.1\\pm0.9\\times 10^{-3}$ Mpc$^{-3}$, taking into account the estimated incompleteness of the sample, an evolution in the number density of a factor of $2.1\\pm0.7$ versus $1.5\\pm0.3 \\times 10^{-3}$ Mpc$^{-3}$ reported by \\citeauthor{Gawiser:2007} (2007) at $z\\simeq3.1$. Our number density is consistent with the value, found by Nilsson et al. (2009) at $z\\simeq2.3$ when we restricted our analysis to objects matching their $\\sim2 \\times$ brighter luminosity limit. At the Ly$\\alpha$ luminosity limit, at which the sample is complete, we calculate a number density of $1.5\\pm0.5\\times 10^{-3}$ Mpc$^{-3}$, that implies an increasing factor of $1.4\\pm0.5$, consistent with that calculated for all the sample. We derive the equivalent width distribution (\\S5.2), representative of the $z\\simeq2.1$ LAE sample in Fig.~\\ref{fig:ew}. As we can see in Fig.~\\ref{fig:ew}a, for log(L(Ly$\\alpha))\\geq42.1$ the sample is unbiased in the sense of rest-frame equivalent width versus Ly$\\alpha$ luminosity. We consider the unbiased brighter half of the sample to build the histogram in Fig.~\\ref{fig:ew}b. Fitting this distribution with an exponential law, this is consistent with that from Gronwall et al. (2007) for the sample at $z\\simeq3.1$ and broader than that found at $z\\simeq2.3$ by Nilsson et al. (2009). In Fig.~\\ref{fig:ew} we associated the value EW~$=400$~ {\\AA} to the objects characterized by an unphysical equivalent width (Appendix, equation \\ref{colorequation}). The objects with $EW_{rest-frame}>250$ present $UB>27$. Most of the objects in the sample with $EW_{rest-frame}<50$ also have log(L(Ly$\\alpha))<42.1$, meaning that their continuum flux boosted them above the narrow-band catalog detection limit. This behavior was less prevalent at $z\\simeq3.1$ by Gronwall et al. (2007), although the 5$\\sigma$ detection limit creates a similar trend, as shown by the blue curve in Fig.~\\ref{fig:ew}a. As it is described in the Appendix, we estimate the observed EW from the observed color $UB$$-$NB3727. Those estimations are in perfect agreement with those obtained from continuum flux density and Ly$\\alpha$ emission line flux. As described in Dayal et al. (2009), the measured EW at the border of the galaxies can be increased by the cooling of collisionally interstellar medium excited HI atoms, while the continuum almost remains unchanged, but intergalactic medium absorption can attenuate Ly$\\alpha$ flux and so decrease the observed EW. The Ly$\\alpha$ luminosity reveals star formation activity inside a galaxy (\\S5.3). Log(L(Ly$\\alpha))=42.1$, the median Ly$\\alpha$ luminosity of our sample, corresponds to SFR(Ly$\\alpha$) = 1.2 M$_{\\odot}$~yr$^{-1}$, as indicated by the dashed-dotted line of Fig.~\\ref{fig:sfr}. In the same figure we observe the range of SFR(UV) values. The median LAE at $z\\simeq2.1$ has a moderate SFR(UV) of $\\sim$ 4 M$_{\\odot}$~yr$^{-1}$. The ratio of $\\sim$1.5 in the values of SFR(UV)/SFR(Ly$\\alpha$), for the unbiased half of the sample, is caused by potentially complex radiative transfer of Ly$\\alpha$ photons in the dusty, possibly clumpy interstellar medium inside the galaxies \\citep{Atek:2009}. Given the overlap in clustering bias it is worth considering whether $z\\simeq2.1$ LAEs could populate the low (stellar) mass tail of continuum-selected star-forming galaxies at $z\\sim2$. We find that the LAE SFR(UV) is 10 times lower than that calculated from UV continuum and H$\\alpha$ line emission by \\citeauthor{Steidel:2004} (2004) for star-forming galaxies at $z\\sim2$. The Kennicutt estimator, used to derive the star formation rate from UV continuum, assumes that the galaxy is at least 10$^7$ yr old with roughly constant SFR. We find (\\S5.4) that 240/250 (96\\%) of $z\\simeq2.1$ LAEs are blue ($B-R$)$<1$, with 73/250 (30\\%) having ($B-R$)$<0$. This is in good agreement with the $z\\simeq3.1$ sample in both criteria. In fact at $z\\simeq3.1$, LAEs with $R<25$ have median color $B-R=0.53$ (\\citealt{Gronwall:2007}). Our result agrees with the findings of \\citet{Nilsson:2009} at $z\\simeq2.3$ in the fraction of LAEs having ($B-R$)$>0$, but their conclusion that most LAEs are ``red'' depended on considering all objects with rising spectra in $f_{\\nu}$ to be red. A reasonable split of galaxies into blue and red is achieved by using ($B-R$)$=1$ as the dividing line, and we suspect that the sample of \\citet{Nilsson:2009} will show similar properties when this is applied. In fact looking at their Fig. 4 and deriving the behavior of the color $B-R$ from the slope $\\beta(B-R)$, we see that their galaxies are essentially blue, based on our definition. The appearance of bimodality in the LAE rest-UV color at $R<25$ is intriguing. The blue branch is presumably dominated by young, dust-free star-forming galaxies, since unobscured (blue) AGN should have been eliminated from our sample due to their X-ray emission. The red branch may contain obscured (dust-reddened) AGN, galaxies with Ly$\\alpha$ emission from recent starbursts but an overall older or dustier stellar population and low-redshift interlopers that will be identified via follow-up spectroscopy. We calculated the evolutionary tracks of galaxies at z$\\sim$2 in the $U-B$ vs $B-R$ plane, generated using the GALAXEV (Bruzual \\& Charlot 2003) code for a constant star formation rate and a range of masses from 25 Myr to 1 Gyr, parameters consistent with LAE SED fits. We see that a 500 Myr old galaxy with dust absorption A$_V=0$ has color $B-R=0$. If it is star forming, the $U-B$ color, corrected for IGM absorption, is also close to zero. Increasing the age the color $B-R$ becomes slightly bigger than 0. However increasing the amount of dust, for example to A$_V\\sim$1, typical for reddened LBG, the star-forming galaxy can assume $B-R$=0.5-0.6. The color $B-R$=1 is achieved by galaxies with significantly more dust than that measured for typical star-forming populations. There is a smaller difference in $B-R$ between young ($<5 \\times10^8$ yr) and old ($>5 \\times10^8$ yr) star-forming populations than the difference produced by the increasing reddening. The observed median($B-R$)~=~0.16 is typical of star-forming galaxies with A$_V$~=~0 and ages of 0.5-1 Gyr or can be consistent with moderate A$_V$ and age 300-500 Myr. Approximately 30 \\% of our sample with negative $B-R$ color is consistent with being very young (age$<$100 Myr) galaxies without dust. We divide in bins of 0.5 magnitude in $R$ and construct Table \\ref{tab:Rchar}, which shows the magnitude range, the median color, EW, SFR from Ly$\\alpha$ and SFR from the UV continuum. These values are transformed into intrinsic ones, taking into account the dust and gas amount (parametrized by stellar E(B-V) and E$_g$(B-V) ) and radiative transfer effects. The median colors lie inside the ``BX\" region except for the faintest bins which have large photometric uncertainties. As expected we observe that the EW values are bigger for the objects that are fainter in the continuum. We calculated $EW_{rest-frame}>250$ for objects with $UB>27$. Statistical fluctuations related to such a faint continua can produce an over-estimation of the equivalent width of these objects. We observe that bright-continuum objects ($UB<24.5$) are also bright in Ly$\\alpha$ luminosity. For low-EW LAEs ($UB-NB3727$ just $\\simeq 0.73$), as expected, the SFR(UV) is significantly bigger than the SFR(Ly$\\alpha$). In the table we also report the standard deviations in the $R$ magnitude bins. In the last column the scatter error is less meaningful, because of the proportionality between $R$ flux density and SFR(UV). It is seen that the scatter is as big as the corresponding quantity. In $B-R$ color it is consistent with that was observed in Fig.~\\ref{fig:BR}. The clustering analysis (\\S5.5) gives information about the LAEs at $z\\simeq2.1$ as a population and their evolution to redshift zero. In Fig.~\\ref{fig:cluster2} we see that LAEs at very high redshift ($z>4$, Ou sign, H09 sign) can evolve into massive LBG at $z\\sim$3 and also reach, in the local Universe, the bias factor typical of elliptical massive galaxies, corresponding to luminosity between 2.5 and 6.0 L* (as indicated by the points in the figure) and halo masses greater than $4.47 \\times 10^{13} $~M$_{\\odot}$. Looking at $z\\sim3$ \\citep{Gawiser:2007} LAEs were observed to be blue galaxies and to be characterized by lower clustering than other galaxy samples at that redshift. They can evolve into star-forming galaxies at $z\\sim2$ (A0 sign) and then to L$^*$ galaxies in the local Universe. At $z\\simeq2.1$ we calculate a bias factor b~$=1.8\\pm$0.3 for our sample of LAEs. This value is consistent with that found using the conditional mass function for progenitors of L$^*$ galaxies in the local Universe. It is also consistent with the value calculated for the subset of ``BX\" galaxies dimmest in $K$-band ($K_{Vega}>21.5$, \\citealt{Adelbergerb:2005}); that is low mass galaxies. This clustering result matches that of dark matter halos with median masses of log(M/M$_{\\odot})= 11.5^{+0.4}_{-0.5} $, which are some of the lowest halo masses probed at this redshift. Our result shows that $z\\sim2$ LAEs could also be descendants of $z\\simeq3.1$ LAEs, depending on how long dust-free star formation occurs and on possible cyclical repetitions of star formation phases. As LAEs at $z\\simeq2.1$ are consistent with being the progenitors of present-day and L$^*$ galaxies at $z=0$, they are likely building blocks of local galaxies with properties similar to the Milky Way and median halo mass $\\geq 2\\times 10^{12}$ M$_{\\odot}$." }, "0910/0910.0266_arXiv.txt": { "abstract": "Much of our understanding of dark matter halos comes from the assumption that the mass-to-light ratio ($\\Upsilon$) of spiral disks is constant. The best way to test this hypothesis is to measure the disk surface mass density directly via the kinematics of old disk stars. To this end, we have used planetary nebulae (PNe) as test particles and have measured the vertical velocity dispersion ($\\sigma_z$) throughout the disks of five nearby, low-inclination spiral galaxies: IC~342, M74 (NGC~628), M83 (NGC~5236), M94 (NGC~4736), and M101 (NGC~5457). By using H{\\sc i} to map galactic rotation and the epicyclic approximation to extract $\\sigma_z$ from the line-of-sight dispersion, we find that, with the lone exception of M101, our disks do have a constant $\\Upsilon$ out to $\\sim$3~optical scale lengths ($h_R$). However, once outside this radius, $\\sigma_z$ stops declining and becomes flat with radius. Possible explanations for this behavior include an increase in the disk mass-to-light ratio, an increase in the importance of the thick disk, and heating of the thin disk by halo substructure. We also find that the disks of early type spirals have higher values of $\\Upsilon$ and are closer to maximal than the disks of later-type spirals, and that the unseen inner halos of these systems are better fit by pseudo-isothermal laws than by NFW models. ", "introduction": "Traditionally, spiral galaxies have three main components: a dynamically hot bulge, a dynamically cold disk (which can sometimes be divided into ``thin'' and ``thick'' components), and a mysterious dark matter halo. However, measuring the mass of each component separately is a challenge, especially in a galaxy's extreme outer regions where the dark matter halo dominates. Even in late-type spirals with minimal bulges, the mass profiles of the dark halos cannot be decoupled from the visible disk mass using rotation curves alone \\citep{bsk04}. Most analyses have therefore relied on the ``maximal disk'' hypothesis, wherein the disk mass-to-light ratio ($\\Upsilon$) is assumed to be constant with radius, and as large as possible to account for the rotation in the inner parts of the galaxy \\citep[\\eg][]{kent86, pw00, sofue03}. Is the disk $\\Upsilon$ really constant and maximal? Perhaps the best way to break the disk-halo degeneracy is to measure the surface mass of a disk directly from the $z$~motions of stars. In systems where the primary stellar motion is circular, the phase-space distribution of stars in the direction perpendicular to the plane is fixed by the disk's gravitational potential \\citep{bt87}. For example, in the Milky Way, star counts and velocity measurements toward the Galactic pole initially suggested that the stellar disk is roughly isothermal \\citep[\\eg][]{oort}. If this were indeed the case, then the integrated face-on stellar velocity dispersion, $\\sigma_z$, would be related to the disk surface mass, $\\Sigma$, by \\begin{equation} \\sigma_z^2(R) = K G \\, \\Sigma(R) h_z, \\label{isothermal} \\end{equation} where $K = 2 \\pi$ is a constant, $G$ is the gravitational constant, and $h_z$ is the scale height of the stars \\citep{vdK88}. Since surveys of edge-on galaxies \\citep[\\eg][]{vdKs82, bm02} imply that $h_z$ is roughly constant with radius, measurements of $\\sigma_z$ can produce estimates of disk mass that are independent of a galaxy's rotation curve. Absorption line studies have shown that, in the inner regions of spirals, $\\sigma_z$ generally follows the exponential law expected from a constant $\\Upsilon$ disk \\citep{b93, g+97, g+00}. However, these surveys were limited by surface brightness, and do not extend more than $\\sim$2~scale lengths from the nucleus. To probe the outer regions of disks, where the dark matter halo is more important, some other technique is required. Planetary nebulae (PNe) are ideal tools for this purpose: PNe are extremely bright in [\\ion{O}{3}], abundant out to $\\gtrsim$5~disk scale lengths, measurable to $\\sim$2~\\kms\\ with fiber-fed spectrographs, and present in all stellar populations with ages between $10^8$ and $10^{10}$~yr. Moreover, these old disk stars are easy to distinguish from other emission-line sources via their distinctive [\\ion{O}{3}]-H$\\alpha$ ratio \\citep{p12}. \\begin{deluxetable*}{cccccccccc} \\tabletypesize{\\scriptsize} \\tablecaption{Program Galaxies\\label{tabBasic}} \\tablewidth{0pt} \\tablehead{&\\colhead{Hubble}&&\\colhead{Distance}&\\colhead{$h_R$}&\\colhead{Disk}&&\\colhead{$v_{max}$}&\\colhead{Number of}&\\colhead{Survey} \\\\ \\colhead{Galaxy} &\\colhead{Type\\tablenotemark{a}} &\\colhead{$i$\\tablenotemark{b}} &\\colhead{(Mpc)\\tablenotemark{c}} &\\colhead{(kpc)\\tablenotemark{d}} &\\colhead{$\\mu_0$\\tablenotemark{e}} &\\colhead{$E(B-V)$\\tablenotemark{f}} &\\colhead{(\\kms)\\tablenotemark{g}} &\\colhead{PN Velocities} &\\colhead{Region} } \\startdata IC~342 &Scd &$31^\\circ$ &$3.5 \\pm 0.3$ &4.24 &19.60 &0.558 &200 &99 &$4\\farcm 8$ \\\\ M74 &Sc &$6.5^\\circ$ &$8.6 \\pm 0.3$ &3.17 &20.20 &0.070 &170 &102 &$4\\farcm 8$ \\\\ M83 &SBc &$24^\\circ$ &$4.8 \\pm 0.1$ &2.45 &19.07 &0.066 &255 &162 &$18\\arcmin$ \\\\ M94 &Sab &$41^\\circ$ &$4.4^{+0.1}_{-0.2}$ &1.22 &18.88 &0.018 &200 &127 &$5\\farcm 8$ \\\\ M101 &Scd &$17^\\circ$ &$7.3 \\pm 0.5$ &4.99 &20.29 &0.009 &250 &60 &$8\\arcmin$ \\enddata \\tablenotetext{a}{From RC3} \\tablenotetext{b}{IC 342: \\citet{cth00}; M74: \\citet{fb+07}; M83: \\citet{p+01}; M94: \\citet{dB+08}; M101: \\citet{ZEH90}} \\tablenotetext{c}{From Paper~I, except for M101, which is from \\citet{M101PNe}; note that this distance has been updated to be consistent with $M^* = -4.47$ \\citep{p12}} \\tablenotetext{d}{IC~342: \\citet[][$I$-band]{wb03}; M74: \\citet[][$R$-band]{m04}; M83: \\citet[][$R$-band]{k+00}; M94: \\citet[][$3.6\\mu$m]{dB+08} and \\citet[][$R$-band]{g+09}; M101: \\citet[][$R$-band]{khc06}} \\tablenotetext{e}{In mag~arcsec$^{-2}$, corrected for galactic inclination. IC~342: \\citet[][$I$-band]{wb03}; M74: \\citet[][$R$-band]{m04}; M83: \\citet[][$R$-band]{k+00}; M94: \\citet[][$R$-band]{g+09}; M101: \\citet[][$R$-band extrapolation]{k+00}} \\tablenotetext{f}{From \\citet{sfd98}} \\tablenotetext{g}{IC 342: \\citet{cth00}; M74: \\citet{fb+07}; M83: \\citet{p+01}; M94: \\citet{dB+08}; M101: \\citet{ZEH90}} \\end{deluxetable*} Planetary nebulae have only recently been used to probe the kinematics of spiral galaxies. The largest study of this kind has been for the Local Group galaxy M31, where the analysis of over 2000 PN velocities has demonstrated a flattening of the disk's line-of-sight velocity dispersion at large galactocentric radii \\citep{M31PNS}. Although the high-inclination of the galaxy precluded any dynamical measurement of disk mass, these data did present evidence for disk flaring at $R \\gtrsim 11$~kpc. The analysis of 140~PN velocities in the moderately inclined ($i \\sim 56^\\circ$) Local Group spiral M33 yielded a similarly surprising result: by assuming a constant scale-height for the galaxy's disk, \\citet{M33PNe} concluded that the disk mass-to-light ratio actually increases with radius, going from $\\Upsilon_V \\sim 0.3$ in the inner regions to $\\Upsilon_V \\sim 2.0$ at $\\sim 9$~kpc. Finally, \\citet{M94PNS} measured the velocities of 67 PNe in the low-inclination ($i \\sim 35^\\circ$) spiral M94, and found a line-of-sight velocity dispersion that declines exponentially over $\\sim$4~disk scale lengths. Papers~I and II of this series \\citep{thesis1, thesis2} presented narrow-band imaging and follow-up spectroscopy of PNe in a sample of large ($r > 7\\arcmin$), nearby ($D < 10$~Mpc), low-inclination ($i \\leq 41^\\circ$) spiral galaxies. Here we use these data to understand the disk kinematics of five galaxies: IC~342, M74, M83, M94, and M101. We introduce our sample in \\S2, and in \\S3, we discuss the problem of the disk scale height, the one parameter of our analysis that cannot be measured. We explore its dependence on galactic radius, its relationship to other disk properties, such as scale length, and alternatives to the isothermal disk assumption. We then begin our analysis in \\S4 by compensating for the fact that none of our nearby galaxies is precisely face-on. We use H{\\sc i} velocity maps to deproject the galactic disks, define the circular velocity as a function of radius, and estimate asymmetric drift. Next, in \\S5 we identify those PNe which likely belong to the galaxies' spheroidal components, and eliminate them from the analysis. We find that very few objects fall into this category: only three in IC~342, four in M74, six in M83, three in M94, and one in M101. Then, in \\S6, we examine the PNe's line-of-sight dispersions without these halo objects. We show that in the inner regions of our galaxies, the line-of-sight velocity dispersion, $\\sigma_{los}$, generally declines as expected for a constant $h_z$, constant $\\Upsilon$ disk. However, once past $\\sim$4~optical scale lengths, $\\sigma_{los}$ flattens out at a value much larger than expected. We consider a few possible explanations for this phenomenon, including the presence of emission-line contaminants in our sample, the effects of internal extinction, and the behavior of large-scale galactic warps. In \\S7, we model the disks' velocity ellipsoids, and extract $\\sigma_z$ from the line-of-sight velocity dispersions. As expected, we find that $\\sigma_R$ and $\\sigma_{\\phi}$ contribute little to the measured dispersion of our nearly face-on disks. In \\S8, we describe the results for each of the five galaxies in our study. We convert our values of $\\sigma_z$ into disk surface masses and calculate disk mass-to-light ratios. Then, in \\S9, we analyze the results from the inner and outer regions of the galaxies. We compute disk and halo rotation curves, compare our disk mass surface densities to those expected from maximal disks, show that pseudo-isothermal cores fit the residual rotation curves better than NFW models, and discuss the physical mechanisms that may produce a constant $\\sigma_z$, including the interaction of thin and thick disks with halo substructure. We conclude in \\S10 by summarizing our results and suggesting future observations that will assist in their interpretation. ", "conclusions": "An analysis of 550 precise PN velocities has yielded kinematic mass estimates throughout the disks of five low-inclination spiral galaxies. In the inner regions of IC~342, M74, and M94, we find that the velocity dispersion perpendicular to the galactic disk exponentially decays with a scale length twice that of the optical light. While our observations of the remaining two galaxies yield ambiguous results (due to a Lindbland resonance in M83 and a poorly sampled velocity field in M101), the data are consistent with expectations from a constant scale height, constant mass-to-light ratio disk. In the later-type (Scd) systems, these disks are clearly sub-maximal, with surface mass densities less than a quarter that needed to reproduce the central rotation curves. In earlier (Sc) galaxies, this discrepancy is smaller, but still present; only the very-early Sab system M94 has evidence for a maximal disk. When we subtract off the disk contribution from the H{\\sc i} rotation curve, we infer the potential of the galaxies' invisible dark halos in all our galaxies except M94. (The uncertainties in the contribution of M94's bright bulge preclude such an analysis in this earliest galaxy in our sample.) In all four cases, these halos are best fit with a simple pseudo-isothermal model; NFW profiles are poor fits to the data, as they always overpredict the residual rotation curve. As in the analysis of low surface-brightness and dwarf galaxies, we find no evidence for central dark matter cusps in our systems. Interestingly, once outside of $\\sim$3~disk scale lengths, the $z$-direction velocity dispersion of galactic disks no longer declines, but instead stays constant with radius. This behavior, coupled with stability arguments, suggests that the disks of spiral galaxies flare substantially at extremely large radii, and may have strongly increasing mass-to-light ratios. Whether this flaring is related to the increased importance of the thick disk, or heating of the thin disk by halo substructure, is unclear at the present time. To further investigate this phenomenon, deep surface photometry of the extreme outer disks of edge-on galaxies is needed. Our kinematic mass estimates for galactic disks are consistent with expectations. The galaxies with the latest Hubble types (IC~342, M101) have the smallest mass-to-light ratios, while earlier systems (M74 and M94) have values of $\\Upsilon$ greater than one. Unfortunately, at this time, no deep, multi-band optical and IR surface photometry exists for any of our galaxies. Once these data are acquired, we will be able to compare our kinematic values for $\\Upsilon$ to stellar mass-to-light ratio estimates from population synthesis models. Such a confrontation will not only produce a better constraint on galactic populations, but also will improve our knowledge of the phase-space distribution of galactic disks." }, "0910/0910.0309.txt": { "abstract": "New $^{12}$CO $J$=4--3 and $^{13}$CO $J$=3--2 observations of the N159 region, an active site of massive star formation in the Large Magellanic Cloud, have been made with the NANTEN2 and ASTE sub-mm telescopes, respectively. The $^{12}$CO $J$=4--3 distribution is separated into three clumps, each associated with N159W, N159E and N159S. These new measurements toward the three clumps are used in coupled calculations of molecular rotational excitation and line radiation transfer, along with other transitions of the $^{12}$CO $J$=1--0, $J$=2--1, $J$=3--2, and $J$=7--6 as well as the isotope transitions of $^{13}$CO $J$=1--0, $J$=2--1, $J$=3--2, and $J$=4--3. The $^{13}$CO $J$=3--2 data are newly taken for the present work. The temperatures and densities are determined to be $\\sim$ 70-80K and $\\sim 3 \\times 10^3$ cm$^{-3}$ in N159W and N159E and $\\sim$ 30K and $\\sim 1.6 \\times 10^3$ cm$^{-3}$ in N159S. Observed $^{12}$CO $J$=2--1 and $^{12}$CO $J$=1--0 intensities toward N159W and N159E are weaker than expected from calculations of uniform temperature and density, suggesting that low-excitation foreground gas causes self-absorption. These results are compared with the star formation activity based on the data of young stellar clusters and H\\emissiontype{II} regions as well as the mid-infrared emission obtained with the Spitzer MIPS. The N159E clump is associated with embedded cluster(s) as observed at 24 $\\micron$ by the Spitzer MIPS and the derived high temperature, 80K, is explained as due to the heating by these sources. The N159E clump is likely responsible for a dark lane in a large H\\emissiontype{II} region by the dust extinction. On the other hand, the N159W clump is associated with embedded clusters mainly toward the eastern edge of the clump only. These clusters show offsets of 20$\\arcsec$ - 40$\\arcsec$ from the $^{12}$CO $J$=4--3 peak and are probably responsible for heating indicated by the derived high temperature, 70 K. The N159W clump exhibits no sign of star formation toward the $^{12}$CO $J$=4--3 peak position and its western region that shows enhanced $R_{4-3/1-0}$ and $R_{3-2/1-0}$ ratios. We therefore suggest that the N159W peak represents a pre-star-cluster core of $\\sim 10^5 M_{\\odot}$ which deserves further detailed studies. We note that recent star formation took place between N159W and N159E as indicated by several star clusters and H\\emissiontype{II} regions, while the natal molecular gas toward the stars have already been dissipated by the ionization and stellar winds of the OB stars. The N159S clump shows little sign of star formation as is consistent with the lower temperature, 30K, and somewhat lower density than N159W and N159E. The N159S clump is also a candidate for future star formation. ", "introduction": "Giant molecular clouds (hereafter GMCs) are the principal sites of star formation and studies of GMCs are important in understanding the evolution of galaxies. A few tens of GMCs in the solar neighborhood such as Orion A have been well studied, while most of the GMCs in the Galaxy are located at more than a few kpc away in the Galactic disk, where contamination by un-related components in the same line of sight seriously limits detailed understanding of GMCs and their associated objects. Recent CO surveys of molecular clouds toward external galaxies in the Local Group have revealed that properties of GMCs such as relations between CO luminosity and line width, the mass range, and the index of mass spectrum are similar among these galaxies \\citep{2007prpl.conf...81B}. This suggests that the properties of GMCs are fairly common among galaxies. The Large Magellanic Cloud (hereafter LMC) is the nearest galaxy to our own at a distance of 50 kpc and is nearly face on, making it an ideal laboratory to observe various properties of GMCs. The small distance enables us to make molecular observations at high spatial resolutions. The LMC also provides a unique opportunity to study molecular clouds and star formation in different environments from the Galaxy; the gas-to-dust ratio is $\\sim 4$ times higher \\citep{1982A&A...107..247K}, and the metal abundance is about $\\sim $3 -- 4 times lower \\citep{2002A&A...396...53R,1984IAUS..108..353D} than those of the Galaxy. Following a low resolution $^{12}$CO $J$=1--0 survey at 140 pc resolution with the Columbia 1.2 m telescope located at CTIO in Chile \\citep{1988ApJ...331L..95C}, high resolution observations in the $^{12}$CO $J$=1--0, $J$=2--1 emission line were made with the SEST 15 m telescope toward some of the molecular clouds and revealed their clumpy structure at 10 pc resolution (e.g., \\cite{1986ApJ...303..186I,1994A&A...291...89J,1996ApJ...472..611C,1997A&AS..122..255K,1998A&A...331..857J,2003A&A...406..817I}). However, these high resolution studies were limited in the spatial coverage, compared with the large angular extent of the LMC, $\\sim 6\\arcdeg \\times 6\\arcdeg$. \\citet{2008ApJS..178...56F} carried out a survey in the $^{12}$CO $J$=1--0 emission line at 40 pc resolution over a $6\\arcdeg \\times 6\\arcdeg$ field in the LMC with the NANTEN 4 m telescope and obtained a complete sample of 270 GMCs (see also for preceding works \\cite{1999PASJ...51..745F,2001PASJ...53L..41F,2001PASJ...53..971M,2001PASJ...53..985Y}). These studies revealed that young stellar clusters whose ages are less than 10 Myr are spatially well correlated with GMCs, and that GMCs are categorized into three types in terms of star formation activity. Type I is starless in the sense that they are not associated with O stars, Type II is associated with small H\\emissiontype{II} regions only, and Type III is associated with huge H\\emissiontype{II} regions and young stellar clusters \\citep{Kawamura09, 2007prpl.conf...81B}. These types are interpreted in terms of evolutionary sequence of GMCs in a timescale of 2-30 Myrs \\citep{1999PASJ...51..745F,2007prpl.conf...81B,Kawamura09}. The $^{12}$CO $J$=1--0 emission line is a probe commonly used to trace molecular clouds because of its low excitation energy ($\\sim$ 5K) and low critical density for collisional excitation ($n_{cr} \\sim 1000$cm$^{-3}$). The $^{12}$CO $J$=1--0 emission alone is however not able to provide physical properties like kinetic temperature and density, the fundamental parameters of GMCs. The high $J$ transitions have higher excitation energies and higher critical densities; e.g., the $^{12}$CO $J$=3--2 transition has the upper state at 33 K and the critical density, $3\\times 10^4 $cm$^{-3}$ \\citep{2005A&A...432..369S}, and the $^{12}$CO $J$=4--3 transition has the upper state at 55 K and the critical density, $1\\times 10^5 $cm$^{-3}$. These sub--mm CO emission lines can selectively trace the sites which may be warmer and denser than the mm CO lines and have a potential to reveal detailed physical properties where star formation is taking place. We are also allowed to make better estimates of temperatures and densities of molecular clouds with a combination of multi--$J$ CO line intensities and molecular excitation analyses such as the Large Velocity Gradient (hereafter LVG) model of molecular line transfer. Recently-developed sub--mm telescopes such as ASTE, NANTEN2 and APEX located at altitudes around 5000m in Atacama, Chile have enabled us to observe higher excited CO transitions at sub--mm wavelengths in superb observational conditions \\citep{2004SPIE.5489..763E,2006IAUSS...1E..21F,2006A&A...454L.115G}. At a lower angular resolution, AST/RO also offered sub-mm observing capability \\citep{2001PASP..113..567S}. \\citet{2000ApJ...545..234B} first detected the $^{12}$CO $J$=4--3 emission line toward the N159 region at 50 pc resolution with the AST/RO telescope. \\citet{2005ApJ...633..210B} later presented estimates of temperatures and densities; $T_{\\rm kin}$ = 20 K, $n({\\rm H}_2)$ = 10$^5$ cm$^{-3}$ for the cold dense component and $T_{\\rm kin}$ = 100 K, $n({\\rm H}_2)$ = 100 cm$^{-3}$ for hot tenuous component toward N159W. Subsequently, \\citet{2008ApJS..175..485M} carried out high resolution $^{12}$CO $J$=3--2 observations of several GMCs at 5 pc resolution with the ASTE 10 m telescope including N159 in the LMC. \\citet{2008ApJS..175..485M} revealed detailed structure of highly excited gas with $T_{\\rm kin} \\sim$ 20 -- 200 K and $n({\\rm H}_2)$ $\\sim$ 10$^3$ -- 10$^4$ cm$^{-3}$. In the present study, we shall focus on the N159 GMC which shows the highest $^{12}$CO $J$=1--0 intensity from the NANTEN survey \\citep{2008ApJS..178...56F}. The N159 GMC is classified as Type III, and includes at least three prominent clumps, N159W, N159E and N159S \\citep{1994A&A...291...89J}. The two molecular clumps in the northern part, N159W and N159E, are associated with massive star clusters whose ages are younger than 10 Myr \\citep{1996ApJS..102...57B} and with huge H\\emissiontype{II} regions \\citep{1956ApJS....2..315H,1976MmRAS..81...89D,1986ApJ...306..130K}. On the other hand, there is no star formation in N159S. In order to better estimate the physical properties of the N159 region, we have carried out new sub-mm observations in the $^{12}$CO $J$=4--3 emission line at a 10 pc spatial resolution with NANTEN2 and in the $^{13}$CO $J$=3--2 emission line at a 5 pc spatial resolution with ASTE. We have also made combined calculations of molecular rotational excitation and line radiative transfer by employing the CO datasets available in N159 and compared the results with the star formation activity. The present paper is organized as follows: Section \\ref{pasj_obs} describes the observations. Sections \\ref{pasj_res} and \\ref{pasj_analysis} show the observational results and data analysis, respectively. We discuss the correlation between highly excited molecular gas and star formation activities in section \\ref{pasj_sf}. Finally, we present a summary in section \\ref{pasj_sum}. ", "conclusions": "\\label{pasj_sum} We carried out $^{12}$CO $J$=4--3 observations of the N159 region with NANTEN2 and $^{13}$CO $J$=3--2 observations toward $^{12}$CO $J$=3--2 peaks with ASTE. The main results are summarized as follows; \\begin{enumerate} \\item The N159 GMC has been resolved into three prominent clumps, N159W, N159E, and N159S in the $^{12}$CO $J$=4--3 emission line. N159W shows the highest intensities among the three in the CO transitions from $J$=1--0 to $J$=7--6. \\item Molecular densities and temperatures have been derived toward the three peaks. Using a LVG analysis involving the $^{12}$CO $J$=4--3 , $^{13}$CO $J$=3--2 emission lines in addition to the other CO lines published, we find that N159W and N159E have temperature of $\\sim$ 70 -- 80 K and density of $\\sim 3 \\times 10^3$ cm$^{-3}$, and that N159S has temperature of $\\sim$ 30 K and density of $\\sim 1.6 \\times 10^3$ cm$^{-3}$. In order to explain lower line intensities than expected, we suggest that $^{12}$CO $J$=1--0 and $J$=2--1 lines may be affected by self-absorption by foreground lower excitation gas. \\item The $^{12}$CO $J$=4--3 distribution is compared with H$\\alpha$ and mid-- to far--infrared emission, a sign of embedded star formation, obtained with the Spitzer SAGE program. N159E shows a good coincidence with a dark lane of H$\\alpha$ and also with a 24 $\\micron$ extended source. On the other hand, N159W is associated with three compact 24 $\\micron$ sources and some small H$\\alpha$ features, although the $^{12}$CO $J$=4--3 peak of N159W and its western part show no sign of star formation. \\item A comparison between N159 and $\\eta$ Car indicates that they show similar star formation activity and we do not see significant difference in physical parameters between these two massive star forming regions. \\end{enumerate} \\bigskip We thank the all members of the NANTEN2 consortium and ASTE team for the operation and persistent efforts to improve the telescopes. This research was supported by the Grant-in-Aid for Nagoya University Global COE Program, \"Quest for Fundamental Principles in the Universe: from Particles to the Solar System and the Cosmos\", from the Ministry of Education, Culture, Sports, Science and Technology of Japan. This work is financially supported in part by a Grant-in-Aid for Scientific Research from the Ministry of Education, Culture, Sports, Science and Technology of Japan (No. 15071203) and from JSPS (Nos. 14102003 and 18684003), and by the JSPS core-to-core program (No. 17004), and the Mitsubishi Foundation. This work is also financially supported in part by the grant SFB 494 of the Deutsche Forschungsgemeinschaft, the Ministerium fur Innovation, Wissenschaft, Forschung und Technologie des Landes Nordrhein-Westfalen and through special grants of the Universitat zu Koln and Universitat Bonn. SAGE research has been funded by NASA/Spitzer grant 1275598 and NASA NAG5-12595. The ASTE project is driven by Nobeyama Radio Observatory (NRO), a branch of National Astronomical Observatory of Japan (NAOJ), in collaboration with University of Chile, and Japanese institutes including University of Tokyo, Nagoya University, Osaka Prefecture University, Ibaraki University., and Hokkaido University. Observations with ASTE were in part carried out remotely from Japan by using NTT's GEMnet2 and its partner R\\&E (Research and Education) networks, which are based on AccessNova collaboration of University of Chile, NTT Laboratories, and NAOJ. A part of this study was financially supported by the MEXT Grant-in-Aid for Scientific Research on Priority Areas No.\\ 15071202." }, "0910/0910.5219_arXiv.txt": { "abstract": "Measurements of X-ray scaling laws are critical for improving cosmological constraints derived with the halo mass function and for understanding the physical processes that govern the heating and cooling of the intracluster medium. In this paper, we use a sample of \\ngroup X-ray selected galaxy groups to investigate the scaling relation between X-ray luminosity ($\\rm{L}_{\\rm X}$) and halo mass ($\\rm{M}_{\\rm 200}$) where $\\rm{M}_{\\rm 200}$ is derived via stacked weak gravitational lensing. This work draws upon a broad array of multi-wavelength COSMOS observations including 1.64 degrees$^2$ of contiguous imaging with the Advanced Camera for Surveys (ACS) to a limiting magnitude of $\\rm{I}_{ \\rm F814W}=26.5$ and deep {\\sl XMM-Newton/Chandra} imaging to a limiting flux of $1.0\\times 10^{-15}$ \\flux in the 0.5-2 keV band. The combined depth of these two data-sets allows us to probe the lensing signals of X-ray detected structures at both higher redshifts and lower masses than previously explored. Weak lensing profiles and halo masses are derived for nine sub-samples, narrowly binned in luminosity and redshift. The COSMOS data alone are well fit by a power law, ${\\rm M}_{\\rm 200} \\propto (\\rm{L}_{\\rm X})^{\\alpha}$, with a slope of \\cosalpha . These results significantly extend the dynamic range for which the halo masses of X-ray selected structures have been measured with weak gravitational lensing. As a result, tight constraints are obtained for the slope of the \\mlx relation. The combination of our group data with previously published cluster data demonstrates that the \\mlx relation is well described by a single power law, \\henkalpha, over two decades in mass, $\\rm{M}_{\\rm 200} \\sim 10^{13.5}$ -- $10^{15.5} \\mass$. These results are inconsistent at the \\sssig\\ level with the self-similar prediction of $\\alpha=0.75$. We examine the redshift dependence of the \\mlx relation and find little evidence for evolution beyond the rate predicted by self-similarity from $z \\sim 0.25$ to $z \\sim 0.8$. ", "introduction": "Groups and clusters of galaxies, formed through the gravitational collapse of massive dark matter halos, are now readily identified up to redshift one and even beyond \\citep[\\eg][]{Stanford:2006,Eisenhardt:2008}. Baryonic tracers such as red-sequence galaxies, typically abundant at the centers of groups and clusters, or X-ray emission from the hot intracluster medium (ICM), have proved to be especially successful in this task \\citep[\\eg][]{Gladders:2005,Koester:2007,Finoguenov:2007}. Nonetheless, these observables only trace the tip of the iceberg given that the vast majority of the underlying mass is in the form of dark matter. The quantification of the total mass (both dark and baryonic) of groups and clusters of galaxies is an important endeavour from both a cosmological and an astronomical standpoint. In particular, several lines of research would benefit from a clearer understanding of the relationship between the total halo mass of groups and clusters and their baryonic tracers. We outline several briefly here \\citep[for a recent review on this subject see][]{Voit:2005}. From a cosmological perspective, the number density of groups and clusters as a function of total mass is of fundamental interest because it is sensitive to both the expansion and growth history of the universe and can be used to constrain cosmological parameters such as $\\Omega_m, \\sigma_8,$ and $\\Omega_{\\Lambda}$ \\citep[\\eg][]{White:1993,Wang:1998a,Haiman:2001,Rozo:2004, Wang:2004a, Bahcall:2004, Rozo:2009}. Furthermore, modifications to the laws of gravity which can be invoked as a possible physical mechanism for acceleration, could imprint telltale signatures in the abundance and dark matter structure of groups and clusters of galaxies \\citep[][]{Rapetti:2008,Schmidt:2008}. Unfortunately, the common difficulty encountered with each of these enquires is that theories and simulations make dark matter based predictions but our most accessible observables (such as richness or X-ray luminosity) are baryonic in nature. It has long been recognized that baryonic observables are subject to complex and poorly understood physical processes that make them imperfect dark matter tracers. For example, X-ray studies discovered early on that the theory of pure gravitational collapse which makes simple predictions for the shape and amplitude of X-ray scaling relations \\citep[also known as the {\\it self-similar model,}][]{Kaiser:1986}, fails to match observations such as the slope and normalization of the \\lx-T relation \\citep[][and references therein]{Voit:2005} implying that other non-gravitational (and still much debated) processes have significantly affected the thermodynamic state of the ICM. Additional heating and cooling mechanisms that are invoked to solve this puzzle lead to different predictions for the shape and redshift evolution of X-ray scaling relations. On the one hand, the fact that X-ray scaling relations deviate from simple models is a plague for cosmologists because there is no straightforward recipe to estimate total halo masses. On the other hand, from an astronomical perspective, the comparison between X-ray observables and total halo mass contains valuable clues about the physical processes that govern galaxy formation and the heating and cooling of the ICM. For all of these reasons, more precise measurements of the mean and scatter in the relationship between total halo mass and various baryonic tracers of groups and clusters of galaxies are highly desirable (\\eg\\ see discussions in \\citet{Voit:2005} and \\citet{Albrecht:2006}). At present, there are five popular methods for detecting groups and clusters of galaxies: a) optical detection via the red-sequence \\citep[\\eg][]{Gladders:2005,Koester:2007}, b) detection via the Sunyaev-Zeldovich (SZ) effect which measures the distortion of the CMB spectrum due to the hot ICM \\citep[\\eg][]{Sunyaev:1970,Sunyaev:1972,Carlstrom:2002,Benson:2004,Staniszewski:2008}, c) detection via X-ray emission \\citep[\\eg][]{Bohringer:2000,Hasinger:2001,Finoguenov:2007, Vikhlinin:2008}, d) spectroscopic identification \\citep[\\eg][]{Gerke:2005,Miller:2005,Knobel:2009}, and e) detection via weak lensing maps \\citep[\\eg][]{Marian:2006,Miyazaki:2007,Massey:2007a}. This last technique is the simplest in terms of the underlying physics and is the only method for which the total halo mass can be directly probed, independently of both the baryons and the dynamical state of the cluster. However, shear maps can only detect the most massive systems ($M>10^{14} M_{\\odot}$) and are limited to moderate redshifts because the lensing weight function peaks mid-way between the source and the observer, with galaxy shapes increasingly difficult to measure at $z>1$. X-ray observations on the other hand, can more simply probe complete samples of groups and clusters, but departures from virial equilibrium and non-thermal pressure components in the ICM can bias X-ray based hydrostatic mass estimates \\citep[\\eg][]{Nagai:2007}. The SZ effect has the attractive property of being redshift independent, and the integrated SZ flux increment, Y, may be less sensitive to the baryon physics of cluster cores \\citep[][]{Motl:2005,Nagai:2006} but mass measurements with the SZ effect face other challenges such as the identification and removal of radio point-sources \\citep[][]{Vale:2006}, sky confusion owing to projection effects \\citep{White:2002a}, and possibly a larger scatter in the Y-M relation than previously estimated due to feedback processes \\citep[][]{Shaw:2008}. Given these considerations, a promising strategy is to employ a robust and efficient cluster detection method \\citep[to which the ultimate solution may be a combination of several techniques such as described in][]{Cohn:2009} and to perform an absolute mass calibration of baryonic tracers via weak gravitational lensing \\citep[][]{Hoekstra:2007a,Rykoff:2008,Berge:2008}. The focus of this paper is to advance these goals by calibrating the slope and amplitude of the \\mlx relation for galaxy groups using cross-correlation weak lensing in the COSMOS survey (also called ``group-galaxy lensing''). Extending weak lensing measurements into the group regime is particularly important in order to extend the dynamic range of weak-lensing-based mass-estimates so as to more accurately determine the slopes of scaling relations. Using the COSMOS sample, we show that X-ray detections span a more complete and wide range of redshift and mass than detections via shear maps. Indeed, high redshift and small structures are challenging to detect directly with shear because of the shape of the lensing weight function (see $\\S$\\ref{group_selection}). Nevertheless, once they have been identified, groups and clusters can be studied via stacking techniques. A notable advantage of this method is that measurements are unaffected by uncorrelated mass along the line-of-sight whereas mass estimates for individually detected clusters are subject to $\\sim 20 \\%$ uncertainties \\citep[][]{Metzler:2001,Hoekstra:2003a,de-Putter:2005}. The associated drawback with stacking is that the intrinsic scatter around the mean relation is difficult to recover. In order to employ the stacked weak lensing technique, tight baryonic tracers of halo mass are highly desirable. The X-ray luminosity of groups and clusters is considered to be a reasonable tracer of halo mass with a logarithmic scatter in the \\mlx relation of roughly $20\\%$ to $30\\%$ \\citep[][]{Stanek:2006, Maughan:2007, Pratt:2008, Rozo:2008, Rykoff:2008, Vikhlinin:2009}. A large fraction of this scatter has been shown to be associated with the presence of cool-cores and simple excision techniques can reduce the scatter to sub-$20\\%$ levels \\citep[][]{Maughan:2007,Pratt:2008}. Although more tightly correlated mass tracers have been identified such as the ${\\rm Y_{X}}$ parameter \\citep[\\eg][]{Kravtsov:2006} -- such indicators require the measurement of an X-ray spectrum which is not possible for most survey data where count rates are low. Our choice of \\lx as a mass proxy reflects that fact that it is a simple X-ray observable, accessible with survey quality data, and the only one that can be easily measured at high redshift. Temperature measurements may be feasible for a small fraction of high redshift objects but cosmological studies that require complete samples of high redshift groups and clusters will need simple mass proxies like ${\\rm{L_{\\rm X}}}$. The details of the \\mlx relation are also important (regardless of the choice of a mass proxy) for determining effective volumes as a function of mass in X-ray flux limited surveys \\citep[][]{Stanek:2006, Vikhlinin:2009}. The layout of this paper is as follows. The data are presented in $\\S$\\ref{cosmos_survey} and the theoretical lensing background is developed in $\\S$\\ref{lensing_theory}. The construction of the group catalog and the lens selection are specified in $\\S$\\ref{group_selection}. Details regarding the adopted form of the \\mlx relation are given in $\\S$\\ref{lx_m_relation}. Our main results are presented in $\\S$\\ref{results} followed by our assessment of the systematic errors in $\\S$\\ref{syst_error}. Finally, we discuss the results and draw up our conclusions in $\\S$\\ref{discussion}. We assume a WMAP5 $\\Lambda$CDM cosmology with $\\Omega_{\\rm m}=0.258$, $\\Omega_\\Lambda=0.742$, $\\Omega_{\\rm b}h^2=0.02273$, $n_{\\rm s}=0.963$, $\\sigma_{8}=0.796$, $H_0=72$ $h_{72}$ km~s$^{-1}$~Mpc$^{-1}$ \\citep[][]{Hinshaw:2009}. All distances are expressed in physical units of $h_{72}^{-1}$ Mpc. X-ray luminosities are expressed in the 0.1-2.4 keV band, rest-frame. The letter M denotes halo mass in general whereas \\m is explicitly defined as $M_{200}\\equiv M($10 keV background yields a more accurate determination of $R$. In this paper, we report results from an observation of the nucleus of NGC 4593 with {\\it Suzaku} in 2007, with the goals of constraining the Fe K emission profile and accurately determining the strength of the Compton reflection hump in order to constrain the geometry of the circumnuclear accreting gas. As demonstrated below, {\\it Suzaku} caught the source at an atypically low 2--10 keV flux level, a factor of almost 4 lower than during the {\\it XMM-Newton} observation. We observe significant changes in the Fe K$\\alpha$ profile between the 2002 {\\it XMM-Newton} and 2007 {\\it Suzaku} observations which may be related to the decrease in continuum flux. We also report evolution in the strength of the soft excess, and, tentatively, the ionized Fe K emission. The rest of this paper is organized as follows: Section 2 describes the observations and data reduction. In Section 3, we present fits to the Fe K emission complex observed with {\\it Suzaku} and compare the results to those for the 2002 {\\it XMM-Newton} observation. In Section 4, we present fits to the 0.3--76 keV broadband {\\it Suzaku} time-averaged spectrum, and again compare the results to those obtained from {\\it XMM-Newton} to investigate long-term spectral variability of the broadband emission components. The results are discussed in Section 5. ", "conclusions": "\\subsection{Summary of Observational Results} We have presented results from a {\\it Suzaku} observation of the nucleus of the Seyfert AGN NGC 4593 in 2007 December, and we compare our spectral fits for both the Fe K bandpass and the broadband X-ray spectrum with those obtained from a 2002 {\\it XMM-Newton} EPIC-pn observation. {\\it Suzaku} caught the source at a relatively low X-ray flux level: the 2--10 keV continuum flux during the {\\it Suzaku} observation was a factor of 3.8 lower. The Fe K$\\alpha$ line intensity has dropped by a factor of 1.7, suggesting that roughly half of the total line flux has responded to the drop in continuum flux. One of our main results is that the Fe K$\\alpha$ line is significantly more narrow in the {\\it Suzaku} observation. Modeling the line as a single Gaussian, we find that the width $\\sigma$ has dropped from $87 \\pm 17$ eV in 2002 to $41^{+12}_{-16}$ eV in 2007. We also modeled the line using a dual-Gaussian model composed of relatively narrow and broad lines. The former dominates the {\\it Suzaku} profile and is assumed to be time-invariant; in the {\\it XMM-Newton} spectrum, both lines are modeled to have roughly equal intensity and the broad component has a width $\\sigma = 177^{+84}_{-52}$ eV. There is highly tentative evidence for the \\ion{Fe}{26} emission line at 6.96 keV to have dropped in intensity from 2002 to 2007, assuming an intrinsically narrow (unresolved) line: in the {\\it Suzaku} observation, the line is not significantly detected ($EW < 17$ eV). In our broadband fits to the 0.3--76 keV spectrum, the primary power-law component, commonly attributed to inverse Comptonization of soft seed photons in a hot corona (e.g., Haardt et al.\\ 1994), was relatively flat, with $\\Gamma$ near 1.65. The Compton reflection component had a relative strength $R \\sim 1.08$. We also model a modest soft excess using both thermal bremsstrahlung emission and thermal Comptonization of soft seed photons, similar to B07, and we obtain similar results. Importantly, we find the soft excess has dropped in flux by a factor of at least $\\sim$20 between the {\\it XMM-Newton} and {\\it Suzaku} observations. We model one zone of absorption along the line of sight, the previously seen highly-ionized (log$\\xi$ $\\sim$ 2.5) zone, with a column density similar to that obtained by McKernan et al.\\ (2003) and Steenbrugge et al.\\ (2003). There is no strong evidence for evolution of the warm absorber between the two observations. A relativistically broadened Fe K$\\alpha$ line was not detected in the {\\it Suzaku} spectrum; Reynolds et al.\\ (2004) demonstrated a similar result in the {\\it XMM-Newton} spectrum. \\subsection{Tracing the Truncated Disk with the Fe K$\\alpha$ Line} % We explore two (model-dependent) scenarios to correlate the changes in Fe line intensity and profile with the observed drop in continuum flux. In the model where a single Gaussian was used to describe the Fe K$\\alpha$ profile, the width $\\sigma$ in the 2007 {\\it Suzaku} observation was $41^{+12}_{-16}$ eV, which corresponds to a FWHM velocity $v_{\\rm FWHM}$ of $4420^{+1290}_{-1730}$ km s$^{-1}$. This is roughly commensurate with the optical broad emission line width: Peterson et al.\\ (2004) reported FWHM H$\\alpha$ and H$\\beta$ line widths of $3399\\pm196$ and $3769\\pm862$ km s$^{-1}$, respectively.\\footnote{Peterson et al.\\ (2004) reported that the continuum-line lag results were poor. The H$\\alpha$ lag was reported as $3.2^{+5.6}_{-4.1}$ lt-days but Peterson et al.\\ (2004) recommended caution. The H$\\beta$ lag was reportedly ``completely unreliable.''} Assuming that the line originates in gas that is in virialized orbit around the black hole, we can estimate the distance $r$ from the black hole to the line-emitting gas. Assuming that the velocity dispersion is related to $v_{\\rm FWHM}$ as $<$$v^2$$>$ = $\\frac{3}{4}v^2_{\\rm FWHM}$ (Netzer 1990), we use $G$$M_{\\rm BH}$ = $r$$v^2$. We use a black hole mass $M_{\\rm BH}$ of $6.6 \\times 10^6 \\Msun$, an estimate based on the relation between $M_{\\rm BH}$ and stellar velocity dispersion in Seyferts (Nelson et al.\\ 2004). The best-fit reverberation mapping estimate from Peterson et al.\\ (2004), $5.4 \\times 10^6 \\Msun$, is consistent with this estimate. We find $r$ = $6.0^{+10.2}_{-2.4} \\times 10^{13}$ m, or $2.3^{+3.9}_{-1.0}$ lt-days. As 1 $R_{\\rm g}$ corresponds to $1 \\times 10^{10}$ m for the black hole mass used, $r$ = $6000^{+10500}_{-2500} R_{\\rm g}$. We cannot of course rule out contribution from an even more narrow Gaussian component originating in even more distant material. In the 2002 {\\it XMM-Newton} observation, the corresponding measured line width (we use our best-fit value of $\\sigma = 87 \\pm 17$ eV) corresponds to a value of $r$ = $1.3^{+0.7}_{-0.4} \\times 10^{13}$ m = $0.50^{+0.30}_{-0.15}$ lt-days, or $1350^{+750}_{-350} R_{\\rm g}$ (see also B07). B07 also used the lack of observed variability in the Fe line flux during the {\\it XMM-Newton} observation to constrain the light-crossing time for the line-emitting gas to be at least 2000 $R_{\\rm g}$. One possible explanation to explain the change in Fe K$\\alpha$ line profile, insofar as it traces the geometrically thin, radiatively efficient disk, is that the innermost radius of the thin disk has increased over 5 years. A common model for accretion flows incorporating truncated thin disks is one where the thin disk transitions to an inner RIAF as the flow crosses a certain transition radius $r_{\\rm t}$ (Esin et al.\\ 1997); a commonly invoked type of RIAF is an advected-dominated accretion flow (ADAF), wherein the disk is optically thin and geometrically thick (e.g., Narayan \\& Yi 1995). The largest width observed for the Fe K$\\alpha$ line thus could indicate $r_{\\rm t}$. In this model, $r_{\\rm t}$ is expected to increase, and more of the inner thin disk evaporates, as the accretion rate relative to Eddington, $\\dot{m}$, decreases in a given object. Supporting evidence for this comes from timing observations of black hole X-ray Binary systems during outburst decay: characteristic temporal frequencies in the power spectral density function (PSD), such as peaks of Lorentzian components and/or quasi-periodic oscillations, migrate towards lower temporal frequencies as $\\dot{m}$ decreases and the source luminosity fades, as the source evolves through the low/hard spectral state into quiescence (e.g., Axelsson et al.\\ 2005, Belloni et al.\\ 2005). In addition, the temperature and flux of the soft, thermally emitted component have been seen to decrease with $\\dot{m}$ in many sources (e.g., Gierli\\'{n}ski, Done \\& Page 2008). However, the predicted relationship between $\\dot{m}$ and $r_{\\rm t}$ remains unclear. Yuan \\& Narayan (2004) empirically derive that compact sources accreting at $\\dot{m}$ near $10^{\\sim-2}$, $10^{\\sim-4}$, and $10^{\\sim-(6-7)}$ may be associated with values of $r_{\\rm t}$ near $10^{\\sim(1-2)}$, $10^{\\sim(2-3)}$, and $10^{\\sim(4-5)}$ $R_{\\rm g}$, respectively. Assuming a 2--10 keV flux in 2002 of $3.9 \\times 10^{-11}$ erg cm$^{-2}$ s$^{-1}$ from {\\it RXTE-PCA} monitoring, a luminosity distance of 41.3 Mpc (following Mould et al.\\ 2000, and using $H_{\\rm o}$ = 70 km s$^{-1}$ Mpc$^{-1}$ and $\\Lambda_{\\rm o}$ = 0.73), the 2--10 keV luminosity $L_{2-10} = 7 \\times 10^{42}$ erg s$^{-1}$. From Marconi et al.\\ (2004), an AGN with this $L_{2-10}$ has a bolometric luminosity $L_{\\rm bol} = 15 L_{2-10} = 1.1 \\times 10^{44}$ erg s$^{-1}$. The accretion relative to Eddington $\\dot{m}$ = $L_{\\rm bol}/L_{\\rm Edd}$ is thus estimated to be 0.15 for the 2002 {\\it XMM-Newton} observation. $\\dot{m}$ during the {\\it Suzaku} observation is thus 0.04. Meanwhile, Lu \\& Wang (2000) have derived $\\dot{m} \\sim 5.5\\%$ from SED fitting. These values of $\\dot{m}$ and our derived value of $r_{\\rm t}$ are not immediately consistent with the rough relation of Yuan \\& Narayan (2004), thus pointing toward models incorporating smaller values of $r_{\\rm t}$ (see below). Furthermore, most low-$\\dot{m}$ sources are radio loud, but NGC 4593 is not strongly radio-loud. Its 5 GHz flux has been measured to near $\\sim$2 mJy (e.g., Schmitt et al. 2001), and its B-band flux is $\\sim$6--16 mJy (e.g., McAlary et al.\\ 1983), so the radio loudness parameter, defined as the ratio of these two values, is $\\lesssim$3. Values $\\geq$10 define a source as radio-loud (Kellermann et al.\\ 1989). A connection between $\\dot{m}$ (proportional to the observed X-ray flux) and $r_{\\rm t}$ in NGC 4593 is thus qualitative only as well as speculative, especially since we have only two model-dependent estimates of $r_{\\rm t}$. There is also the question of whether the inner portions of a thin disk in AGN can evaporate and/or become radiatively inefficient on timescales of only a few years. As the accretion disks of BH XRBs are thought to evolve on timescales of at least hours to days, the corresponding timescales in NGC 4593 (black hole mass a factor of $10^{\\gtrsim5}$ higher) would be decades to centuries. On the other hand, Marscher et al.\\ (2002) interpreted rapid (duration of $\\sim$ a couple weeks) dips in the X-ray light curve of the radio-loud Seyfert 3C~120 as periods when the inner portion of the disk evaporated, each event leading to ejection of material into the relativistic jet and a corresponding radio flare about a month later. The total Fe K$\\alpha$ line intensity decreased from $4.79 \\pm 0.48 \\times 10^{-5}$ (2002) to $2.93 \\pm 0.23 \\times 10^{-5}$ (2007) ph cm$^{-2}$ s$^{-1}$, a factor of 1.7, while the observed 4--10 keV flux decreased from $2.47\\times 10^{-11}$ (2002) to $0.69 \\times 10^{-11}$ (2007) erg cm$^{-2}$ s$^{-1}$, a factor of 3.6, i.e., roughly half of the total line flux has responded to continuum decrease. Modeling the Fe K$\\alpha$ line with a dual-Gaussian model attempted to separate the variable and non-variable emission components; in this model, a non-variable, narrow component is detected in both observations, and dominates the total line flux in the {\\it Suzaku} spectrum, while a broader component is detected only in the {\\it XMM-Newton} spectrum. The best-fit width $\\sigma$ of the broad line was $177^{+84}_{-52}$ eV, corresponding to FWHM velocity of $19100^{+9100}_{-5600}$ km s$^{-1}$, implying $r = 3.2^{+3.3}_{-1.7} \\times 10^{12}$ m = $0.12^{+0.12}_{-0.06}$ lt-days, or $330^{+330}_{-180} R_{\\rm g}$. This estimate is inconsistent with B07's estimate of $\\sim$2000 $R_{\\rm g}$ based on the invariance of the Fe K$\\alpha$ line during the {\\it XMM-Newton} observation, but it is still consistent with the presence of a truncated thin disk ($r_{\\rm t} > 150 R_{\\rm g}$). $r_{\\rm t}$ could be invariant from 2002 to 2007, but an annular region on the thin disk spanning from $\\sim 300 R_{\\rm g}$ to 1000--5000 $R_{\\rm g}$ (outer radius obviously speculative) could be responding to the drop in illuminating continuum flux. If the drop in \\ion{Fe}{26} flux is real, then that line could also originate in this region. However, it is not clear in this model why material at $\\gtrsim 6000 R_{\\rm g}$ (yielding the narrow line component), contributing roughly half of the total line intensity in 2002, has not responded to the drop in continuum flux, as it is well within a week's light-travel time. The width of the narrow line had been fixed at $\\sigma=41$ eV in our modeling of the {\\it XMM-Newton} profile, but contributions from a more narrow component likely cannot be ruled out. Such distant material could be responding to a previous higher continuum flux. The 2--10 keV {\\it RXTE}-PCA monitoring light curve in fact showed a higher, more average flux level until roughly 300 days before the {\\it Suzaku} observation (Figure 7). A final possibility that does not require evolution in $r_{\\rm t}$ is that the inner disk may have become be too ionized to transmit an Fe line. In the context of models with a hot, ionized skin (Nayakshin, Kazanas \\& Kallman 2000), a disk illuminated by a power-law continuum with a photon index near 1.6, similar to that in the {\\it Suzaku} observation, yields extremely weak Fe line emission. Of course, both profile models are likely oversimplifications. The community could thus benefit from an X-ray observatory with $\\sim$ few eV resolution combined with a large effective area near 6 keV to resolve the various components of the Fe K$\\alpha$ core as a function of time and/or continuum flux level, if multiple components do indeed exist, as well as resolve the \\ion{Fe}{26} line. \\subsection{The Compton Reflection Component} % A Compton shoulder was not significantly detected in either the {\\it Suzaku} or {\\it XMM-Newton} spectra; we find upper limits to Compton shoulder emission (first-scattering) of $\\sim$23$\\%$ of the core. It is thus not obvious from this limit alone whether bulk of the Fe K$\\alpha$ line originates in Compton-thick material, especially the degree to which the strength of the Compton shoulder depends on the geometry of the material. However, no relativistically broadened Fe K$\\alpha$ line has been confirmed in NGC 4593, but we confirm from the broadband {\\it Suzaku} spectrum the presence of a Compton reflection hump which thus must correspond to (at least some fraction of) the Fe K$\\alpha$ core emission. We can investigate if the measured strength $R$ of the Compton reflection hump, $1.08 \\pm 0.20$ (statistical uncertainty only; $\\pm$ 0.35 including the systematic uncertainty), can correspond to the observed Fe K$\\alpha$ line $EW$ of 255$\\pm$19 eV. Following George \\& Fabian (1991), one expects an $EW$ of 135 eV (using the abundances of Lodders 2003) to correspond to $R$ = 1 for an semi-infinite optically thick slab illuminated by a power-law continuum with $\\Gamma$ = 1.7 and assuming solar abundances and an inclination angle of 30$\\degr$ relative to the observer's line of sight. The observed values of $R$ and $EW$ are thus consistent if the Fe abundance relative to solar, $Z_{\\rm Fe}$, is about 1.7, which is not unreasonable. For a truncated disk, this could be explained by having the Comptonizing corona consist of numerous flares lying in a sandwich-like geometry just above/below the thin disk (e.g., Haardt et al.\\ 1994), such that the disk spans 2$\\pi$ sr of the sky as seen by each X-ray flare. However, very distant (pc-scale), Compton-thick material lying out of the line of sight, which cannot be ruled out as contributing to the observed Fe K emission profile, could also contribute to the total Compton reflection strength. However, if the thin disk is truncated, then a semi-infinite slab may not be an appropriate geometry, especially if the central X-ray source is not immediately close to the reflecting disk. The $EW$ of a truncated disk will be lower, but will depend on the location of the illuminating X-ray source. If we assume that the illuminating X-ray source is located on the disk symmetry axis a height $h$ above the disk, we can use the $EW$ to constrain $h$. George \\& Fabian (1991, their Fig.\\ 15) demonstrate that the reflected flux is dominated by the region of the disk with $r/h \\sim$ 1--2. For a truncated thin disk with $r_{\\rm t}$ = several thousand $R_{\\rm g}$, $h$ must also be $\\sim$ several thousand $R_{\\rm g}$ above the black hole. A configuration in which the X-ray corona is associated with the base of a jet along the symmetry axis may thus be applicable, e.g., Markoff, Nowak \\& Wilms (2005). NGC 4593, like many Seyferts, is known to host a pc-scale radio component (size of $<$15 pc; Schmitt et al.\\ 2001). \\subsection{Spectral Variability of the Broadband Components} The primary-law component in Seyfert X-ray spectra is usually attributed to inverse Comptonization of soft seed photons. In an ADAF flow, thermal Comptonization is expected to dominate the X-ray emission unless the accretion rate is extremely low, in which case thermal bremsstrahlung emission dominates the X-ray spectrum (Narayan et al.\\ 1998, Narayan 2005). One of our main results is that while the primary power-law component has dropped in flux by a factor of almost 4, the soft excess has dropped in flux by a factor of $\\sim$20 between 2002 and 2007, ruling out an origin for the soft excess with a size greater than 5 lt-years. This drop is likely linked to the decrease in the primary X-ray power-law, i.e., it may be either a cause of an effect of it. We explored two phenomenological models for the soft excess, bremsstrahlung and thermal Comptonization. In the presence of an ADAF flow, one can expect thermal bremsstrahlung emission with a temperature of $10^{9-11}$ K (Narayan \\& Yi 1995), but such temperatures are higher by over 2 orders of magnitude compared to the temperature derived from our model fits and by B07. We also modeled the soft excess as inverse Comptonization of soft seed photons with an assumed input temperature of 50 eV by an optically thin corona with a temperature $k_{\\rm B}T \\sim 50$ keV, again obtaining similar results to B07. The location of such a process is not clear, though it could occur in the ionized skin of the thin disk (e.g., Magdziarz et al.\\ 1998, Janiuk et al.\\ 2001), or at the base of an outflowing jet. If both the soft excess and hard X-ray power-law components originate via Comptonization of disk seed photons, a decrease in the intensity of those soft seed photons (e.g., from an increase in $r_{\\rm t}$) between 2002 and 2007 could yield a correlated drop in both component's flux. Another possibility is that the optical depth of the Comptonizing components may have changed." }, "0910/0910.4359_arXiv.txt": { "abstract": "% We present results of the investigation of the nature of double periodic variables (DPVs). We have selected a sample of Galactic eclipsing DPVs for a multiwavelength photometric study aimed to reveal their nature. The short orbital periodicity and the cyclic variability are decoupled and separately investigated. Shapes of orbital light curves are consistent with semi-detached binaries. The amplitude of the long cycle is always larger in redder bandpasses. ", "introduction": "Double Periodic Variables (DPVs) is a group of binary stars with orbital period between 1 and 16~d, characterized by additional long cyclic variability in the range of 50-600~d, correlated with the orbital period (Mennickent et al.~2003, Mennickent \\& Ko{\\l}aczkowski 2009a). These stars were discovered during the search for Be stars in the Magellanic Clouds based on the OGLE variable stars catalogue. A careful survey of the OGLE and MACHO photometry, allowed to detect this kind of variability in about a hundred stars in the Magellanic Clouds. The first results obtained for one LMC DPV star, OGLE05155332-6925581, were recently published (Mennickent et al.~2008). Investigating the ASAS photometric data we have found eleven Galactic DPVs. The light curves of four Galactic DPVs are shown in Fig.1. After the discovery of this new group of variable stars, we have initiated a multiwavelength observing campaign. ", "conclusions": "The analysis of the available data, suggests that DPVs are binaries in a mass exchange evolution stage. Probably the primary is rotating at the critical velocity and matter cumulates in a circumprimary disc. Disc size increases until certain instability ejects mass outside the binary system. As a result, in DPVs we observe the long-term variability (Mennickent \\& Ko{\\l}aczkowski 2009abc)." }, "0910/0910.3786_arXiv.txt": { "abstract": "{ It is usually assumed that the ellipticity power spectrum measured in weak lensing observations can be expressed as an integral over the underlying matter power spectrum. This is true at order ${\\mathcal O}(\\Phi^2)$ in the gravitational potential. We extend the standard calculation, constructing all corrections to order ${\\mathcal O}(\\Phi^4)$. There are four types of corrections: corrections to the lensing shear due to multiple-deflections; corrections due to the fact that shape distortions probe the reduced shear $\\gamma/(1-\\kappa)$ rather than the shear itself; corrections associated with the non-linear conversion of reduced shear to mean ellipticity; and corrections due to the fact that observational galaxy selection and shear measurement is based on galaxy brightnesses and sizes which have been (de)magnified by lensing. We show how the previously considered corrections to the shear power spectrum correspond to terms in our analysis, and highlight new terms that were not previously identified. All correction terms are given explicitly as integrals over the matter power spectrum, bispectrum, and trispectrum, and are numerically evaluated for the case of sources at $z=1$. We find agreement with previous works for the ${\\mathcal O}(\\Phi^3)$ terms. We find that for ambitious future surveys, the ${\\mathcal O}(\\Phi^4)$ terms affect the power spectrum at the $\\sim 1-5\\sigma$ level; they will thus need to be accounted for, but are unlikely to represent a serious difficulty for weak lensing as a cosmological probe. } ", "introduction": "Cosmic shear, the distortion of light from distant galaxies by the tidal gravitational field of the intervening large scale structure, is an excellent tool to probe the matter distribution in the universe. The statistics of the image distortions are related to the statistical properties of the large scale matter distribution and can thereby be used to constrain cosmology. Current results already demonstrate the power of cosmic shear observations at constraining the clustering amplitude $\\sigma_8$ and the matter density $\\Omega_{\\mr m}$ \\citep[e.g.,][]{Cosmos,Tim,CFHTLS, Fu}. Furthermore, cosmic shear provides an ideal tool to study dark energy through measuring the evolution of non-linear structure with large future surveys (DES\\footnote{www.darkenergysurvey.org}, LSST\\footnote{www.lsst.org},JDEM\\footnote{http://jdem.gsfc.nasa.gov}, Euclid\\footnote{http://sci.esa.int/science-e/www/object/index.cfm?fobjectid=42266}). These upcoming large weak lensing experiments will limit the statistical uncertainties to the percent level. In order to extract cosmological information from these cosmic shear experiments, the increased data quality needs to be accompanied by a thorough treatment of systematic errors. On the observational side, this requires accurate information on the redshift distribution of source galaxies \\citep{MHH_photoz} and precise measurements of galaxy shapes which correct for observational systematics such as pixelization, noise, blurring by seeing and a spatially variable point spread function \\citep[see][]{STEP2,GREAT08}. On the theoretical side, astrophysical contaminants, like source lens clustering \\citep{BvWM97,Schneider02}, intrinsic alignment \\citep{King_ic} and the correlation between the gravitational shear and intrinsic ellipticities of galaxies \\citep{HS04, King_il,JS08,Zhang08,JS09}, need to be understood and removed. The prediction of lensing observables also requires precise models of the non-linear matter power spectrum and models for the relation between lensing distortion and large scale matter distribution which go beyond linear theory. While N-body simulations may predict the non-linear dark matter power spectrum with percent level accuracy in the near future \\citep{Coyote1, Coyote2}, the effect of baryons, which is a significant contamination to the weak lensing signal above $l\\sim 2000$ \\citep{Jing06, Rudd08}, is more difficult to account for and is the subject of ongoing work. In this paper, we consider corrections to the relation between the observed lensing power spectra and the non-linear matter density field. In the regime of weak lensing, the observed galaxy ellipticities ($e_I$) are an estimator of the reduced shear $g_I = \\gamma_I/(1-\\kappa)$, \\be \\ensav{e_I} = C\\frac{\\gamma_I}{1-\\kappa}\\; , \\label{eq:e} \\ee where $C$ is a constant which depends on the type of ellipticity estimator \\citep[e.g.][]{SS95, SS97} and the properties of the galaxy population under consideration, $\\gamma_I$ is a component of the shear, $\\kappa$ is the convergence, and the subscript $I$ refers to the two components of the ellipticity/shear \\citep[see e.g. ][for more details]{BS01}. The two-point statistics of the measured ellipticities are simply related to the reduced shear power spectrum. \\citet{CH} have calculated the shear power spectrum to fourth order in the gravitational potential. For the reduced shear power spectrum there exists an approximation to third order in the gravitational potential \\citep{DSW}. \\citet{Shapiro08} has demonstrated that on angular scales relevant for dark energy parameter estimates the difference between shear and reduced shear power spectra is at the percent level and ignoring these corrections will noticeably bias dark energy parameters inferred from future weak lensing surveys. \\citet{FS_lens} introduced another type of corrections, termed \\emph{lensing bias}, which has a comparable effect on the shear power spectrum as the reduced shear correction: Observationally, shear is only estimated from those galaxies which are bright enough and large enough to be identified and to measure their shape. This introduces cuts based on observed brightness and observed size, both of which are (de)magnified by lensing \\citep[e.g.][]{BTP_mag, Jain_size}, and will thus bias the sampling of the cosmic shear field. In the following we complete the calculation of the reduced shear power spectrum to fourth order in the gravitational potential to include multiple deflections and to account for the effects of lensing bias and the non-linear conversion between ellipticity and reduced shear. We consider all lensing-related effects through ${\\mathcal O}(\\Phi^4)$, but do {\\em not} include effects associated with the sources (source clustering and intrinsic alignment corrections). This paper is organized as follows: We describe our technique for calculating higher order lensing distortions and power spectra in Sect.~\\ref{sec:method}. Derivations of the different types of corrections to the shear and reduced shear power spectra are given in Sect.~\\ref{sec:shear} through Sect.~\\ref{sec:lensbias}. We quantify the impact of these corrections on future surveys in Sect.~\\ref{sec:impact} and discuss our results in Sect.~\\ref{sec:discussion}. ", "conclusions": "\\label{sec:discussion} We have calculated the reduced shear power spectra perturbatively to fourth order in the gravitational potential, accounting for the differences between shear and reduced shear, relaxing the Born approximation, and including lens-lens coupling in the calculation of shear and convergence. The full set of corrections to the reduced shear power spectra are given in Table \\ref{tab:cg} (E-mode) and Table \\ref{tab:cgB} (B-mode). The ellipticity power spectrum contains additional contributions, Eq.~(\\ref{eq:NLcorr}), which arises from the non-linearity of the shear estimator and depends on the specific definition of ellipticity used, and Eq.~(\\ref{eq:LBE}) which is caused by lensing bias. Through order $\\Phi^4$, this is the full set of corrections to the power spectrum arising from the lensing process itself. All corrections have been derived within the Limber approximation, and the analysis of ``12\" type multiple-deflection corrections is left for future work. Other corrections associated with the source galaxy population, such as source clustering and intrinsic alignments, are not treated in this paper. We find that, depending on the properties of the source galaxy population and on the type of shear estimator used, these corrections will be at the $\\sim 1-5\\sigma$ level, and thus should be included in the analysis of future precision cosmology weak lensing experiments. That said, we caution that there are other areas in which the theory of weak lensing needs work if it is to meet ambitious future goals. Current fitting formula of the non-linear dark matter power spectrum have an accuracy of about $10\\%$ at arcminute scales \\citep{sm03} and the uncertainty exceeds $30\\%$ for $l>10000$ \\citep{HH09}, due to this difficulty in modeling the non-linear gravitational clustering angular scales of $l> 3000$ are likely to be excluded from parameter fits to cosmic shear measurements. Utilizing near-future $N$-body simulations it will become possible to determine the non-linear dark matter power spectrum with percent level accuracy \\citep[e.g.,][]{Coyote1, Coyote2}. However, this does not account for the effect of baryons, which will likely be important at halo scales and depend critically on the details of baryonic processes (cooling, feedback) involved. Baryons in dark matter halos which are able to cool modify the structure of the dark matter halo through adiabatic contraction \\citep{Blumenthal86,Gnedin04}, causing deviations of the inner halo profile from the simple NFW form and changing the halo mass - halo concentration relation \\citep[e.g.][]{Rudd08, Pedrosa09}. The latter can be constrained though galaxy-galaxy lensing \\citep{MandelbaumGG}, or could be internally self-calibrated in a weak lensing survey via its preferential effect on the small-scale power spectrum \\citep{Zentner08}. Baryons in the intergalactic medium may make up about $10$\\% of the mass in the universe, and if their distribution on Mpc scales has been strongly affected by non-gravitational processes then they could pose a problem for precise calculation of the matter power spectrum \\citep[see][for an extreme and probably unrealistic example]{LG06}. Given these uncertainties in modeling the non-linear matter distribution and that all the corrections derived in this paper are integrals over the non-linear matter power spectrum, bispectrum and trispectrum, we refrain from calculating $\\mathcal O(\\Phi^5)$ and higher corrections. We expect that the corrections derived in this paper are sufficient to model the perturbative relation between the non-linear matter distribution and the lensing distortion in weak lensing surveys for the forseeable future. \\begin{acknowledgement} E.K. and C.H. are supported by the US National Science Foundation under AST-0807337 and the US Department of Energy under DE-FG03-02-ER40701. C.H. is supported by the Alfred P. Sloan Foundation. We thank Wayne Hu, Fabian Schmidt, Peter Schneider and Chaz Shapiro for useful discussions. \\end{acknowledgement}" }, "0910/0910.1260_arXiv.txt": { "abstract": "We report new estimates of the time delays in the quadruple gravitationally lensed quasar PG1115+080, obtained from the monitoring data in filter $R$ with the 1.5-m telescope at the Maidanak Mountain (Uzbekistan, Central Asia) in 2004-2006. The time delays are 16.4 days between images C and B, and 12 days between C and A1+A2, with image C being leading for both pairs. The only known estimates of the time delays in PG1115 are those based on observations by Schechter et al. (1997) -- 23.7 and 9.4 days between images C and B, C and A1+A2, respectively, as calculated by Schechter et al., and 25 and 13.3 days as revised by Barkana (1997) for the same image components with the use of another method. The new values of time delays in PG 1115+080 may be expected to provide larger estimates of the Hubble constant thus decreasing a diversity between the $H_0$ estimates taken from gravitationally lensed quasars and with other methods. ", "introduction": "As was first suggested by Refsdal (1964), gravitationally lensed quasars can potentially provide an estimate of the Hubble constant $H_0$ independent of any intermediate distance ladder. This can be made from measurements of the time delays between the quasar intrinsic brightness variations seen in different quasar images. The value of $H_0$ can be obtained then (within the adopted cosmological model) from the observed geometry of the system, with the known lens and source redshifts, and with the use of physically validated model of mass distribution in the lensing galaxy. Since a phenomenon of gravitational lensing is controlled by the surface density of the total matter (dark plus luminous), it provides a unique possibility both to determine the value of $H_0$ and to probe the dark matter abundance in lensing galaxies and along the light paths in the medium between the quasar and observer. By now the time delays have been measured in about 20 gravitationally lensed quasars resulting in the values of $H_0$, different for different quasars, while remaining noticeably less than the most recent estimate of $H_0$ obtained in the HST Hubble Constant Key Project with the use of Cepheids - $H_0=72\\pm8$ km s$^{-1}$ Mpc$^{-1}$ (Freedman et al. 2001). This discrepancy is large enough and, if the Hubble Constant is really a universal constant, needs to be explained. A detailed analysis of the problem of divergent $H_0$ estimates inherent in the time delay method and the ways to solve it can be found, e.g., in Keeton \\& Kochanek (1997); Kochanek (2002); Kochanek \\& Schechter (2004); Schechter (2005). The quadruply lensed quasars, and PG1115+080 in particular, are known to better suit for determining the $H_0$ value as compared to the two-image lenses since they provide more observational constraints to fit the lens model. The PG1115 source quasar with a redshift of $z_S=1.722$ is lensed by a galaxy with $z_G=0.31$ (Henry \\& Heasley 1986; Christian et al. 1987; Tonry 1998), which forms four quasar images, with an image pair A1 and A2 bracketing the critical curve very close to each other. It is the second gravitationally lensed quasar discovered over a quarter of century ago, at first as a tripple quasar (Weymann et al. 1980). Hege et al. (1980) were the first to resolve the brightest image component into two images separated by 0.48 arcsec. Further observations (Young et al. 1981; Vanderriest et al. 1986; Christian et al. 1987; Kristian et al. 1993) have provided positions of quasar images and information about the lensing object, which allowed to build a model of the system (e.g., Keeton, Kochanek \\& Seljak 1997). In particular, Keeton \\& Kochanek (1997) have shown that the observed quasar image positions and fluxes and the galaxy position can be fit well by an ellipsoidal galaxy with an external shear rather than by a just ellipsoidal galaxy or a circular galaxy with an external shear. They noted that a group of nearby galaxies detected by Young et al. (1981) could provide the needed external shear. The problem of determining the Hubble constant from the time delay lenses is known to suffer from the so-called central concentration degeneracy, which means that, given the measured time delay values, the estimates of the Hubble constant turn out to be strongly model-dependent. In particular, models with more centrally concentrated mass distribution (lower dark matter content) provide higher values of $H_0$, more consistent with the results of the local $H_0$ measurements than those with lower mass concentration towards the centre (more dark matter). \\begin{figure*} \\resizebox{\\hsize}{!}{\\includegraphics{Fig1.eps}} \\caption{The light curves of PG 1115+080 A1, A2, B, C from observations in filter $R$ with the 1.5-m telescope of the Maidanak Observatory in 2004, 2005 and 2006.} \\end{figure*} The time delays in PG 1115+080 were determined for the first time by Schechter et al. (1997) to be $23.7\\pm 3.4$ days between B and C, and $9.4\\pm 3.4$ days between A1+A2 and C (image C is leading). Barkana (1997) re-analyzed their data using another algorithm and reported $25^{+3.3}_{-3.8}$ days for the time delay between B and C. There are also estimates of the time delays between images A1 and A2 made from the {\\it Chandra} and {\\it XMM-Newton} X-ray Observatories data (Dai X, et al. 2001, Chartas et al. 2004). As is predicted by all lens models, it does not exceed a small fraction of the day, and equals $0.16\\pm0.02$ days (Chandra data) or $0.149\\pm0.006$ days (XMM-Newton data). Determination of the time delays has generated a flow of models for the system, (Schechter et al. 1997; Keeton \\& Kochanek 1997; Courbin et al. 1997; Impey et al. 1998; Saha \\& Wiiliams 2001; Kochanek, Keeton \\& McLeod 2001; Zhao \\& Pronk 2001; Chiba 2002; Treu \\& Koopmans 2002; Yoo et al. 2005, 2006; Pooley et al. 2006; Miranda \\& Jetzer 2007), all illustrating how strongly the estimated value of $H_0$ depends on the adopted mass profiles of the lens galaxy for the given values of time delays. ", "conclusions": "" }, "0910/0910.5597_arXiv.txt": { "abstract": "We apply the Standardized Candle Method (SCM) for Type II Plateau supernovae (SNe II-P), which relates the velocity of the ejecta of a SN to its luminosity during the plateau, to 15 SNe II-P discovered over the three season run of the Sloan Digital Sky Survey - II Supernova Survey. The redshifts of these SNe - $0.027 < z < 0.144$ - cover a range hitherto sparsely sampled in the literature; in particular, our SNe II-P sample contains nearly as many SNe in the Hubble flow ($z > 0.01$) as all of the current literature on the SCM combined. We find that the SDSS SNe have a very small intrinsic \\emph{I}-band dispersion ($0.22$ mag), which can be attributed to selection effects. When the SCM is applied to the combined SDSS-plus-literature set of SNe II-P, the dispersion increases to $0.29$ mag, larger than the scatter for either set of SNe separately. We show that the standardization cannot be further improved by eliminating SNe with positive plateau decline rates, as proposed in \\citet{P09}. We thoroughly examine all potential systematic effects and conclude that for the SCM to be useful for cosmology, the methods currently used to determine the \\fe\\ velocity at day 50 must be improved, and spectral templates able to encompass the intrinsic variations of Type II-P SNe will be needed. ", "introduction": "Type Ia supernovae (SNe Ia) are standardized candles; although their peak magnitudes can vary by up to $\\sim 2.5$ mag, empirical relations between the shape of their light curves and peak magnitude result in distance measurements that can be accurate to $\\sim$7\\% \\citep{Phillips,JRK,Guy}. This standardization has proven to be invaluable, as observations of SNe Ia led to the discovery of the accelerating expansion of the universe \\citep{Riess98,Perl99}. Dedicated surveys have now observed over a thousand of these objects in the past several years over a wide range of redshifts, lowering statistical uncertainties with the volume of their discoveries (For high redshift SN programs, see \\citealp{Riess04,Riess07,astier06,WV07,miknaitis07,Kessler09,Frieman08}; for low redshift, see \\citealp{LOSS,CSP,SNF,CFA}). As a result, selection effects and other potential sources of bias are of increasing importance to improving the constraints on the dark energy equation of state. In order to fully utilize future surveys that will discover very large numbers of SNe (DES\\footnote{https://www.darkenergysurvey.org}; LSST\\footnote{http://www.lsst.org/lsst}), these systematic uncertainties will have to be reduced. Type II Plateau Supernovae (SNe II-P) can also be used as distance indicators, though thus far with less precision and to much lower distances than SNe Ia. SNe II-P are objects that, observationally, are classified as Type II SNe by the presence of hydrogen absorption in their spectra, and as `Plateau' based on the slow decay of their early light curves \\citep{barbon79}. Distance measurements using the Expanding Photosphere Method \\citep[EPM; see][]{EPM1,EPM2,EPM3} and the more recent, related Spectral Expanding Atmosphere Method \\citep[SEAM; see][]{SEAM1,SEAM2}, which are based on the modeling of the SN atmosphere, have been shown to recover distances to 10\\% precision \\citep{SEAM3}. A different approach has been taken by \\citet[hereafter HP02]{HP02}, who have shown that SNe II-P magnitudes can also be standardized empirically, despite being far more heterogeneous \\citep[peak luminosity dispersion of more than 5 mags; see][]{Patat,Pastorello} than SNe Ia. The theoretically based methods take advantage of the relative simplicity of modeling hydrogen, which dominates the atmosphere of SNe II-P, as compared to the intermediate mass elements that make up the ejecta of SNe Ia. SNe II-P also differ from their brighter brethren in a couple of other advantageous ways. Over the past decade, observations in archival pre-SNe II-P images have found red supergiants in the mass range $~7-16\\, M_{\\odot}$ \\citep{IIP1,IIP2,IIP3,IIP4,IIP5,IIP6,IIP7,IIP8}, and recently it has been shown that in at least one of these cases, the supposed progenitor is no longer present after the SN II-P has dimmed \\citep{IIP9}. So, compared to SNe Ia, where there is still uncertainty over the progenitor system (or systems), SNe II-P explosions are better understood. Also, since SNe II-P have only been found in late-type galaxies - unlike SNe Ia, which are found in both late- and early-type galaxies - it is likely that biases from environmental effects will have a smaller effect on distance measurements for SNe II-P than SNe Ia. Thus the differences between the two types of SNe and their methods of standardization will result in different systematic effects, allowing SNe II-P to serve as a useful check on SNe Ia measurements. The standardization of the SNe II-P luminosity was introduced by HP02, who found that the luminosity and the expansion velocity at the photosphere are correlated when the SN is in its plateau phase; they use 50 rest-frame days post explosion as a convenient reference epoch to compare across SNe (this epoch is late enough to ensure the SN has entered the plateau phase, while still before it leaves the plateau). The origin of this relationship is strongly grounded physically; a more luminous supernova's hydrogen recombination front will be at a greater distance from the core than in a less luminous SN, and thus the velocity of matter at the photosphere will be greater. Using the velocity of the ejecta as measured from the minimum of the \\feII\\ absorption feature and a reddening correction based on the color of the SN when its light curve falls off the plateau, they found that the scatter in the Hubble diagram for \\emph{V} and \\emph{I} magnitudes drops from 0.95 and 0.80 mag, respectively, to 0.39 and 0.29 mag. When they then restricted their sample to the 8 SNe with $z > 0.01$ in order to reduce the effect of peculiar velocities on host galaxy redshift, the scatter of their sample dropped even further, to 0.20 mag in \\emph{V} and 0.21 mag in \\emph{I}. This technique has come to be known as the Standardized Candle Method (SCM). \\citet[hereafter N06]{N06} improved upon the HP02 method in several ways that address its limitations for application to SNe at cosmological distances. The HP02 host galaxy extinction correction is replaced by the rest-frame \\VmI\\ color at day 50, which can be obtained with less extensive late time photometric monitoring. To allow spectroscopic follow-up programs to be more flexible, they also derive an observational relationship between the velocity of the \\fe\\ line at epochs spanning days 9-75 to the velocity at day 50, which is crucial for any realistic spectroscopic follow-up program. They apply the standardization method to 5 high-redshift ($0.13 < z < 0.29$) SNe obtained as part of the Supernova Legacy Survey \\citep[SNLS;][]{astier06}, as well as one at $z = 0.019$ followed up by the Caltech Core-Collapse Program \\citep[CCCP;][]{CCCP} (SN 1999gi, which was added to the dataset of SNe II-P by \\citet{H03}, is also included). The result of their SNLS-only fit differed strongly from that of the low-redshift SNe in HP02, which was attributed to a number of factors, including a small sample size, differences in data analysis techniques, and observational biases. Their data did confirm, however, that the relation between ejecta velocity and magnitude could be used to reduce the intrinsic dispersion in the sample. \\citet{Olivares} applied the SCM to 37 SNe II-P, for which they use 30 days before the end of the plateau as the common epoch of reference. They find the Hubble Diagram dispersion to be remarkably similar whether one uses \\emph{B}-, \\emph{V}-, or \\emph{I}-band magnitudes (0.28, 0.31, and 0.32 mags, respectively), and raise the possibility that the magnitude-velocity relation may be quadratic instead of linear. The most recent work on SNe II-P standardization was done by \\citet[hereafter P09]{P09}, which presented new data on 19 low-redshift (all located at $z < 0.03$) SNe II-P with multi-epoch spectra. These SNe were discovered as part of the Lick Observatory SN Search \\citep[LOSS;][]{LOSS} by the Katzman Automatic Imaging Telescope (KAIT). The most significant advancement of the SCM presented in P09 is the introduction of a robust technique for determining the velocity of the ejecta by cross-correlating the observed spectra with high-quality SNe II-P templates contained in SNID \\citep[SuperNova Identification Code; see][]{SNID1,SNID2}. As mentioned in P09, the velocity of the weak, broad \\fe\\ line is difficult to measure without introducing systematic offsets that hinder comparisons across multiple data sets. Using the SCM, they found a significantly larger \\emph{I}-band scatter (0.38 mag) in their Hubble Diagram than either HP02 or N06 (including all objects from both previous works), but showed that by removing all SNe with a positive plateau decline rate (defined as a decrease in brightness in \\emph{I}-band magnitude between days 10 and 50) from their analysis, the dispersion of this `culled' sample was only 0.22 mag, or $\\sim 10\\%$ in distance. In this work we apply the SCM to both the 15 new SNe II-P from the SDSS-II Supernova Survey and the combined set of our SNe plus those in the literature. Our sample fills in a redshift range, $0.027 < z < 0.144$, that has very few other objects. We discuss our observations in \\S\\ref{sec-obs}, and present light curves for these SNe and an additional 19 spectroscopically confirmed SNe II-P, also discovered during the SDSS-II Supernova Survey, for which we do not have sufficient data to include in our SCM sample. We also discuss how we determine which SNe are part of the SCM sample. In Section \\S\\ref{sec-phot} we explain in detail our process for determining K- and S-corrections, and explore how these are affected by the uncertainty in the explosion date and the choice of spectral template. We analyze the spectra of our SNe in \\S\\ref{sec-spec}, and obtain their ejecta velocity at day 50. In \\S\\ref{sec-cosm} we present the results from combining our observed SNe with those of HP02, N06, and P09, and applying the SCM to a combined sample of 49 SNe II-P. We discuss our results and their interpretations in \\S\\ref{sec-results}, and conclude with an eye towards what will be needed for future SNe II-P campaigns in \\S\\ref{sec-dis}. ", "conclusions": "\\label{sec-dis} We have presented photometric and spectroscopic data for 15 SNe II-P from the SDSS-II Supernova Survey. These SNe span a range of redshifts not covered by other surveys, connecting the local sample of HP02 and P09 to the high-$z$ SNLS sample in N06. We propagate through our photometry reduction the uncertainty in the time of explosion in a way that accurately represents the state of knowledge regarding the explosion epoch. We also perform K- and S-corrections for all of our SNe and describe the process in detail. The best fit parameters for the SCM are updated with our new data, and are found to change significantly from those presented in P09. The difference can be primarily attributed to the SDSS SNe being intrinsically brighter than the SNe in the `culled' sample of P09 (due to understood selection effects), without displaying the corresponding increase in ejecta velocity predicted by the SCM parameters of P09. We show that the act of warping a SN II-P spectral template to the observed colors, while not completely removing systematic uncertainties, reduces them to such a level that they are unlikely to account for the magnitude of our offset. The difficulty in accurately measuring the velocity of the broad \\fe\\ absorption features and the uncertainty involved in extrapolating velocities at early epochs to later times are likely to be the dominant sources of systematic uncertainty in our analysis. The low $\\alpha$ value we derive for the SDSS-only sample is due to both the small scatter in observed magnitudes and the uncertainty in our velocity determination. Despite the selection effects and measurement uncertainties that limit the power of the SDSS SNe II-P to constrain the SCM parameters, we observe the primary relationship defining the SCM, which is that brighter SNe have faster ejecta, at $\\approx 2\\sigma$. The primary objective of the SDSS-II Supernova Survey was to identify and characterize SNe Ia; as a result, the observing strategy was not ideal for SNe II-P studies. A future survey program that would focus on the SCM would greatly benefit from possessing the following qualities: 1) Longer continuous survey duration. For the final six weeks of such a survey, any newly discovered SNe will not have photometry sufficiently far into their plateau to allow extrapolation of their magnitudes at rest-frame day 50. In addition, no SN can have its explosion date pinned down unless it happens after the first survey observation in its field. Since each season of the SDSS-II SNS was three months long, no more than $50 \\%$ of the survey time was useful for discovering new SNe II-P. Doubling the length of such a survey would nearly triple the number of SNe discovered for which the SCM can be used. 2) Deeper, later spectroscopic observations. When a new SN is discovered in a survey dedicated to SNe II-P, an initial quick spectrum is not needed. A SN II-P will have a distinct light curve from a SN Ia, and thus valuable spectroscopic resources will not be used to follow up objects of no interest. Since the velocity of the ejecta at day 50 is what is required, a spectrum at early times is not required, and observations can target day 50. As we have shown, the systematic uncertainties associated with early-time spectra require a thorough study. Also, each object must be observed for a sufficient time to obtain a high quality S/N spectrum. Accurate measurements of \\feII\\ absorption lines are key to determining the strength of the $\\alpha$ parameter; quality over quantity should be the mantra for these followups. 3) Template database must be extended. To perform both K-corrections and S-corrections, an accurate representation of the true spectral energy distribution is vital. Ideally, the spectrum for each SN in such a survey would have sufficient wavelength coverage to allow corrections for all bands of interest to be computed from it directly. As this situation is unlikely, complementary spectral observations that cover the near IR should allow for templates to be constructed that range from at least $4000 - 12000$\\AA\\, (or redder still, if the wavelengths in the lowest energy filter are redshifted out of this range at the targeted $z$). The amount of uncertainty in the day 50 magnitude results that is due to relying on a limited number of SNe II-P spectral templates will only be known when more extensive templates are available. When these advances can be made, we will be able to determine whether the SCM and SNe II-P can place meaningful, independent constraints on cosmological parameters." }, "0910/0910.5845_arXiv.txt": { "abstract": "We review three Li problems. First, the Li problem in the Sun, for which some previous studies have argued that it may be Li-poor compared to other Suns. Second, we discuss the Li problem in planet hosting stars, which are claimed to be Li-poor when compared to field stars. Third, we discuss the cosmological Li problem, i.e. the discrepancy between the Li abundance in metal-poor stars (Spite plateau stars) and the predictions from standard Big Bang Nucleosynthesis. In all three cases we find that the ``problems'' are naturally explained by non-standard mixing in stars. ", "introduction": "As illustrated in Fig. 1, the present day solar Li abundance \\citep[$A_{\\rm Li}$\\footnote{$A_X = \\log (N_X/N_H) + 12$} = 1.05;][]{asp09} is much lower than the meteoritic Li abundance \\citep[$A_{\\rm Li}$ = 3.26;][]{asp09}. This large depletion in the observed solar Li abundance by a factor of 160 relative to the primordial solar system composition remains one of the most serious challenges of standard solar models, which, as illustrated in Fig. 1, destroy only a minor amount (0.06 dex) of Li \\citep[e.g.][]{dan84}. It is important to note that modern models using updated OPAL opacities predict too much lithium destruction during the pre-main sequence \\citep[e.g.][]{pia02,dan03}, at variance with the observations of Li in solar-mass stars in young open clusters (see e.g. Fig. 6). When the new low solar abundances \\citep{asp09} are used the problem is reduced \\citep{ses06}, but in general classical models have problems dealing with pre-main sequence convection, so other parameters are invoked to reduce the efficient lithium destruction during the pre-main sequence \\citep[e.g.][]{ven98}. \\subsection{Comparison to solar analogs} A comparison between the Sun and solar analogs of one solar mass and solar metallicity \\citep{lam04} shows the Sun to be ``lithium-poor'' by a factor of 10. This apparent peculiarity in the solar Li abundance, led \\cite{lam04} to suggest that the Sun may be of dubious value for calibrating non-standard models of Li depletion. However, Pasquini et al. (1994) have shown that there are solar-type stars that have a Li abundance as low as in the Sun. Nevertheless, \\cite{pas94} comparison sample of solar-type stars span a wide range in stellar parameters (5400 K $<$ \\teff $<$ 6100 K, 3.6 $<$ log $g$ $<$ 4.6, -1.6 $<$ [Fe/H] $<$ +0.2) and therefore they are not representative of one-solar-mass solar analogs. In Fig. 2, we restrict the comparison sample of \\cite{pas94} to only solar analogs within $\\pm$200K in solar effective temperature, $\\pm$0.3 dex of the solar surface gravity and $\\pm$0.3 dex of the solar metallicity. As we can see, these solar analogs seem to cluster in two groups, one with very high lithium abundances of $A_{\\rm Li}$ $\\sim$ 2.4, i.e., 20 times higher than solar, and the other group with Li abundances as low as solar. Why are there no solar analogs with intermediate Li abundances? Why the number of solar analogs with low Li abundances is much lower than the number of analogs with high Li abundances? Could this be the reason why \\cite{lam04} only found stars with high Li abundances around one solar mass? The lack of stars with intermediate Li abundances in the Pasquini et al. (1994) solar analog sample, and the lack of stars with both intermediate and low Li abundance around one solar mass in the \\cite{lam04} sample are probably telling us that both samples have biases, perhaps due to a selection of stars mostly in one or two evolutionary stages, e.g., mainly young stars, which are known to have high Li abundances. The recent work by \\cite{pas08} for solar analogs and solar twins in the solar-age open cluster M67, shows that solar twins (M67 stars around solar effective temperature) have Li abundance as low as solar, but stars 100 or 200 K hotter span a broad range in Li abundances. So, it is important that the temperature scale of the comparison sample is accurate; otherwise offsets of about 100 K may introduce a bias in the comparison between the Sun and stars. \\subsection{Comparison to solar twins} Solar twins, stars with stellar parameters very similar to the Sun, are ideal targets to see if the Sun is normal (or not) in its Li abundance. Being so similar to the Sun, it is possible to obtain reliable stellar parameters, and provided the Sun and the twins are analyzed (and observed) in a consistent way, the temperature scale is accurate. Furthermore, being selected due to their similarity in colors and luminosity to the Sun, they should span a range of ages very close to solar, avoiding thus potential biases in the selection of comparison stars in only one evolutionary stage very different to the present Sun. Solar twins have been searched for a long time, and although interesting solar twin candidates like 16 Cyg B (HD 186427) were identified in the past, detailed analysis showed that they were significantly different to the Sun \\citep[see review by][]{cay96}. When the first close solar twin (18 Sco) was found \\citep{por97}, it seemed to have a Li abundance near solar, but much better data \\citep[e.g.][]{mel06} showed that its Li abundance is actually three times higher than solar. One solar twin is certainly not an acceptable number for a comparison between the Sun and stars, so a large survey of solar twins was urgently needed. The two largest recent efforts for finding field solar twin stars are being undertaking by the group of Y. Takeda \\citep[e.g.][]{tak07} and by our group \\citep{mel06,mel07,mel09a,ram09}. Importantly, whenever possible, we are obtaining very high S/N for our sample stars, because otherwise only upper limits can be obtained for their Li abundances. Indeed, as shown by \\cite{tak07} in their Fig. 12, their data with S/N $\\sim$150 can only estimate upper limits for stars with $A_{\\rm Li}$ $<$ 1.5, i.e., they can only reliably determine Li abundances when they are three times higher than solar. Our solar twin survey has been performed mainly with the 2.7m telescope at McDonald observatory in the North and with the 6.5m Magellan Clay telescope at Las Campanas observatory in the South. We have also obtained some Keck+HIRES data in the North and VLT+UVES and HARPS data in the South. Our data has been taken at R = 60,000-110,000 and achieving S/N = 200-1000. The first pilot data set taken at Keck resulted in the discovery of the second best solar twin, HD 98618 \\citep{mel06}, about a decade after the discovery of the first solar twin 18 Sco. HD 98618 seems to be a solar twin as good as 18 Sco, and, as this twin, it has also a Li abundance three times higher than solar. Learning from the experience of our pilot Keck observations, we improved our criteria to select the best solar twins, empirically adjusting our \\teff scale \\citep{ram05} for an apparent zero-point problem \\citep{cas09}. This is probably the reason why our first solar twin run at McDonald was very successful. Besides confirming the solar twin nature of 18 Sco and HD 98618, we identified two additional solar twins, HIP 56948 and HIP 73815 \\citep{mel07}, both with a low Li abundance similar to solar. HIP56948 remains to this date the star that most closely resembles the Sun, with a \\teff similar to solar within 10 K, as recently confirmed by \\cite{tak09} using Subaru+HDS observations. The year 2007 was very prolific for solar twin studies, besides the twins found by our group, \\cite{tak07} reported the discovery of the fifth solar twin, HIP 110963. Interestingly, this solar twin has a high Li abundance of $A_{\\rm Li}$ = 1.7. These five solar twins were starting to fill the Li desert seen in Fig. 2. The year 2009 has been even better, with many more solar twins found using our McDonald data \\citep{ram09} and Magellan+MIKE observations \\citep{mel09a}, which together have found more than 30 stars very similar to the Sun. Thus, we are starting to have a large sample of solar twins for meaningful statistics, in particular for addressing the long-standing question on chemical peculiarities in the Sun \\citep{mel09a,ram09,gus09}. In Fig. 3 we show the Li abundance for solar twins and solar analogs that have been analyzed for Li in our survey. Most stars from the Magellan run have already been analyzed, but the McDonald Li analysis is just starting. Open circles show detections and filled circles upper limits. The Sun is also shown for comparison. The first important point to note in this plot is that, unlike the lack of stars with both intermediate and low Li in the \\cite{lam04} sample, and the lack of stars with intermediate Li abundances in the \\cite{pas94} sample, our sample nicely covers a broad range of Li abundances 0.6 $<$ $A_{\\rm Li}$ $<$ 2.4, meaning probably that our sample is not affected by any significant selection bias. In Fig. 4 we show the Li abundances vs. \\teff of our solar twin and solar analog sample restricted to stars with mass within $\\pm$3\\% solar and [Fe/H] within $\\pm$0.1 dex solar. The Sun does not look peculiar on this plot. One star with relatively high Li abundance stands out. As shown in Fig. 5, where Li is plotted as a function of age, the high Li abundance of this star is due to its young age. This plot has a smaller number of stars than Figs. 3-4 because here, besides the constraint to $\\pm$3\\% in mass and $\\pm$0.1 dex in [Fe/H], we show only stars for which we could determine reliable ages ($\\ge$ 2.5 sigma). This figure definitely shows that the solar Li abundance is not abnormal, at least not for a solar-metallicity solar-age one-solar-mass star, which at about 4.6 Gyr has already depleted a significant fraction of its original Li abundance. A comparison of our solar twin results to one-solar-mass stars in solar metallicity ($\\pm$0.15 dex) open clusters \\citep[selected from the sample of][and including the results from Pasquini et al. 2008]{ses05} is shown in Fig. 6. Again, the agreement is excellent, and reinforces a strong correlation between Li depletion and age for one-solar-mass stars. Non-standard models \\citep[e.g.][]{mon00,chat05,xio09,don09} can reproduce the observed data, as shown in Figs. 7 and 8. We are in the process of obtaining better Li abundances and ages for our sample stars, which can potentially constrain the range of initial rotational velocities of our solar twins. Based on the results shown above, we conclude that the solar Li abundance is not peculiar but a product of depletion due to non-standard mixing which affect both the Sun and the solar twins. ", "conclusions": "" }, "0910/0910.0717_arXiv.txt": { "abstract": "Although today there are many observational methods, Type Ia supernovae (SNIa) is still one of the most powerful tools to probe the mysterious dark energy (DE). The most recent SNIa datasets are the 307 SNIa ``Union'' dataset \\cite{kow08} and the 397 SNIa ``Constitution'' dataset \\cite{hic09}. In a recent work \\cite{wei10}, Wei pointed out that both Union and Constitution datasets are in tension with the observations of cosmic microwave background (CMB) and baryon acoustic oscillation (BAO), and suggested that two truncated versions of Union and Constitution datasets, namely ``UnionT'' and ``ConstitutionT'', should be used to constrain various DE models. But in \\cite{wei10}, only the $\\Lambda$CDM model is used to select the outliers from the Union and the Constitution dataset. In principle, since different DE models may select different outliers, the truncation procedure should be performed for each different DE model. In the present work, by performing the truncation procedure of \\cite{wei10} for 10 different models, we demonstrate that the impact of different models is negligible, and the approach adopted in \\cite{wei10} is valid. Moreover, by using the 4 SNIa datasets mentioned above, as well as the observations of CMB and BAO, we perform best-fit analysis on the 10 models. It is found that: (1) For each DE model, the truncated SNIa datasets not only greatly reduce $\\chi _{min}^{2}$ and $\\chi _{min}^{2}/dof$, but also remove the tension between SNIa data and other cosmological observations. (2) The CMB data is very helpful to break the degeneracy among different parameters, and plays a very important role in distinguishing different DE models. (3) The current observational data are still too limited to distinguish all DE models. ", "introduction": "Observations of Type Ia supernovae (SNIa) \\cite{rie98,per99,ton03,kno03,rie04a}, cosmic microwave background (CMB) \\cite{ben03,spe03,spe07,pag07,hin07,kom09} and large scale structure (LSS) \\cite{teg04a,teg04b,teg06} all indicate the existence of dark energy (DE) driving the current accelerating expansion of the universe. The most obvious theoretical candidate of DE is the cosmological constant $\\Lambda$, but it is plagued with the fine-tuning problem and the coincidence problem \\cite{wein89,wein00,sah00,car01,pee03,pad03,cop06}. There are also many dynamical DE models, such as quintessence \\cite{pee88,rat88,rat88,zla99}, phantom \\cite{cal02,car03}, $k$-essence \\cite{arm99,chi00,arm01}, CPL \\cite{che01,lin03}, tachyon \\cite{pad02,bag03}, hessence \\cite{wei05a,wei05b}, Chaplygin gas \\cite{kam01}, generalized Chaplygin gas \\cite{bto02}, holographic \\cite{li04,hua05a,hua05b,li08,li09a,zhax10}, agegraphic \\cite{wei08a,wei08b}, holographic Ricci \\cite{gao09}, Yang-Mills condensate \\cite{zhay07a,zhay07b,wans08a,wans08b}, etc. Although numerous theoretical models have been proposed in the past decade, the nature of DE still remains a mystery. In recent years, the numerical study of DE, i.e. utilizing cosmological observations to constrain DE models, has become one of the most active fields in the modern cosmology \\cite{alb06,wany00,wany04,wany07,hut03,hut05,hua04,zhax04,zhax05,cha06,zhax07a,zhax07b,zhax09,wanb05,wanb06,ma09a,ma09b,wei07,wei08c,wei09}. Although today there are many observational methods, SNIa is still one of the most powerful tools to probe the mysterious DE. In the past decade, many SNIa datasets, such as Gold04 \\cite{rie04b}, Gold06 \\cite{rie07}, SNLS \\cite{ast07}, ESSENCE \\cite{woo07}, Davis \\cite{dav07}, have been released, while the number and quality of SNIa have continually increased. The most recent SNIa datasets are ``Union'' \\cite{kow08} and ``Constitution'' \\cite{hic09}, and they have been widely used in the literature \\cite{shaf09,li09b,qi09,hua09,wu09,chen09,gong10a,gong10b,san09,mor09,wans10}. However, these SNIa datasets are not always consistent with other types of cosmological observations, and are even in tension with other SNIa samples. For examples, in \\cite{jas05,jas06,ness05}, the Gold04 dataset was shown to be inconsistent with the SNLS dataset: the SNLS dataset favors the $\\Lambda$CDM model, while the Gold04 dataset favors the dynamical DE model. In \\cite{ness07}, by comparing the maximum likelihood fits of the CPL parameters ($w_0$, $w_1$) given by different SNIa samples, Nesseris and Perivolaropoulos found that the Gold06 dataset is also in $2\\sigma$ tension with the SNLS dataset. Moreover, they also investigated how to remove this tension. The method is simple. First, they fitted the $\\Lambda$CDM model to the whole 182 SNIa in the Gold06 dataset, and obtained the best-fit parameter value of $\\Lambda$CDM model (for SNIa data only). Then, they calculated the relative deviation to the best-fit $\\Lambda$CDM prediction, $|\\mu_{obs}-\\mu_{\\Lambda CDM}|/\\sigma_{obs}$, for all the 182 data points. Here $\\mu_{obs}$ is the observational value of distance modulus, $\\mu_{\\Lambda CDM}$ is the theoretical value of distance modulus given by the best-fit $\\Lambda$CDM model, and $\\sigma_{obs}$ is the 1$\\sigma$ error of distance modulus. By searching the SNIa samples satisfying $|\\mu_{obs}-\\mu_{\\Lambda CDM}|/\\sigma_{obs}>1.8$, they isolated six SNIa that are mostly responsible for the tension. Further, by using the random truncation method, they demonstrated that these 6 SNIa are systematically different from the Gold06 dataset. In a recent work, by comparing the maximum likelihood fits of the CPL parameters ($w_0$, $w_1$), Wei \\cite{wei10} pointed out that both Union and Constitution dataset are also in tension with the observations of CMB and baryon acoustic oscillation (BAO). Moreover, he also investigated how to remove these tensions. By using the method of truncation of \\cite{ness07}, Wei found out the main sources that are responsible for the tensions: for the Union set, there are 21 SNIa differing from the best-fit $\\Lambda$CDM prediction beyond $1.9\\sigma$ (i.e. $|\\mu_{obs}-\\mu_{\\Lambda CDM}|/\\sigma_{obs}>1.9$); and for the Constitution set, there are 34 SNIa differing from the best-fit $\\Lambda$CDM prediction beyond $1.9\\sigma$. The specific limit of truncation (i.e. $1.9\\sigma$) is chosen based on two considerations: first, the tension between SNIa samples and other observations can be completely removed; second, the number of usable SNIa can be preserved as much as possible (see \\cite{wei10} for details). By subtracting these outliers from the Union dataset and the Constitution dataset, respectively, two new SNIa datasets, ``UnionT'' and ``ConstitutionT'' (``T'' stands for ``truncated''), were obtained. Further, Wei argued that the UnionT and the ConstitutionT datasets are fully consistent with the other cosmological observations, and should be used to constrain various DE models. But in \\cite{wei10}, only the $\\Lambda$CDM model is used to select the outliers from the Union and the Constitution dataset. Since different DE models may select different outliers, one may doubt whether the approach adopted in \\cite{wei10} is valid. In principle, the truncation procedure should be performed for each different DE model. Only if the impact of different models is negligible, one can conclude that the conclusion of Wei is correct. So in the present work, we shall consider 10 different models. The truncation procedure will be performed for all these 10 models, and the corresponding cosmological consequences will be explored. This paper is organized as follows: In Section 2, we briefly describe 10 theoretical models considered in this work. In Section 3, we present the method of data analysis, as well as the SNIa datasets we used in this paper. By performing the truncation procedure of \\cite{wei10} for 10 different models, we demonstrate that the approach adopted in \\cite{wei10} is valid. In Section 4, we show the data fitting results of 10 models, and present the corresponding conclusions. Section 5 is a short summary. In this work, we assume today's scale factor $a_{0}=1$, so the redshift $z$ satisfies $z=a^{-1}-1$; the subscript ``0'' always indicates the present value of the corresponding quantity, and the unit with $c=\\hbar=1$ is used. ", "conclusions": "We will present the results of data fitting in this section. Using the Union, the UnionT, the Constitution and the ConstitutionT dataset, respectively, we list the $\\chi_{min}^{2}$ and the $\\chi_{min}^{2}/dof$ for those 10 models, in table \\ref{table1}, table \\ref{table2}, table \\ref{table3}, and table \\ref{table4}. Here ``SNIa'' means only SNIa data is used in the analysis, ``SNIa+BAO'' means both SNIa data and BAO data are used, ``SNIa+CMB'' means both SNIa data and CMN data are used, and ``SNIa+BAO+CMB'' means all these three types of observational data are taken into account. The main conclusions are summarized as follows. \\begin{table} \\caption{The $\\chi_{min}^{2}$ and the $\\chi_{min}^{2}/dof$ (in the Parentheses) for the 10 models, where the Union dataset is used.} \\begin{center} \\label{table1} \\begin{tabular}{ccccc} \\hline\\hline ~~~Model~~~ & ~~~SNIa~~~ & ~~~SNIa+BAO~~~ & ~~~SNIa+CMB~~~ & ~~~SNIa+BAO+CMB~~~ \\\\ \\hline\\hline ~~~$\\Lambda$CDM~~~ & ~~~$311.936~(1.019)$~~~ & ~~~$313.205~(1.020)$~~~ & ~~~$312.424~(1.018)$~~~ & ~~~$313.594~(1.018)$~~~ \\\\ \\hline ~~~DGP~~~ & ~~~$313.026~(1.023)$~~~ & ~~~$314.319~(1.024)$~~~ & ~~~$339.280~(1.105)$~~~ & ~~~$341.584~(1.109)$~~~ \\\\ \\hline ~~~ADE~~~ & ~~~$313.536~(1.025)$~~~ & ~~~$314.838~(1.026)$~~~ & ~~~$327.216~(1.066)$~~~ & ~~~$329.252~(1.069)$~~~ \\\\ \\hline ~~~XCDM~~~ & ~~~$310.682~(1.019)$~~~ & ~~~$311.966~(1.020)$~~~ & ~~~$312.224~(1.020)$~~~ & ~~~$313.456~(1.021)$~~~ \\\\ \\hline ~~~CG~~~ & ~~~$310.434~(1.018)$~~~ & ~~~$311.628~(1.018)$~~~ & ~~~$311.921~(1.019)$~~~ & ~~~$313.193~(1.020)$~~~ \\\\ \\hline ~~~HDE~~~ & ~~~$310.827~(1.019)$~~~ & ~~~$312.149~(1.020)$~~~ & ~~~$311.218~(1.017)$~~~ & ~~~$312.481~(1.018)$~~~ \\\\ \\hline ~~~RDE~~~ & ~~~$310.682~(1.019)$~~~ & ~~~$311.966~(1.020)$~~~ & ~~~$312.323~(1.021)$~~~ & ~~~$313.801~(1.022)$~~~ \\\\ \\hline ~~~LP~~~ & ~~~$309.984~(1.020)$~~~ & ~~~$310.744~(1.019)$~~~ & ~~~$310.691~(1.019)$~~~ & ~~~$311.978~(1.020)$~~~ \\\\ \\hline ~~~CPL~~~ & ~~~$310.091~(1.020)$~~~ & ~~~$310.896~(1.019)$~~~ & ~~~$310.906~(1.019)$~~~ & ~~~$312.258~(1.020)$~~~ \\\\ \\hline ~~~GCG~~~ & ~~~$310.405~(1.021)$~~~ & ~~~$311.401~(1.021)$~~~ & ~~~$311.925~(1.023)$~~~ & ~~~$313.325~(1.024)$~~~ \\\\ \\hline\\hline \\end{tabular} \\end{center} \\end{table} \\begin{table} \\caption{The $\\chi_{min}^{2}$ and the $\\chi_{min}^{2}/dof$ (in the Parentheses) for the 10 models, where the UnionT dataset is used.} \\begin{center} \\label{table2} \\begin{tabular}{ccccc} \\hline\\hline ~~~Model~~~ & ~~~SNIa~~~ & ~~~SNIa+BAO~~~ & ~~~SNIa+CMB~~~ & ~~~SNIa+BAO+CMB~~~ \\\\ \\hline\\hline ~~~$\\Lambda$CDM~~~ & ~~~$204.568~(0.718)$~~~ & ~~~$205.945~(0.720)$~~~ & ~~~$205.575~(0.719)$~~~ & ~~~$206.794~(0.721)$~~~ \\\\ \\hline ~~~DGP~~~ & ~~~$205.234~(0.720)$~~~ & ~~~$206.634~(0.722)$~~~ & ~~~$226.987~(0.794)$~~~ & ~~~$229.388~(0.799)$~~~ \\\\ \\hline ~~~ADE~~~ & ~~~$205.560~(0.721)$~~~ & ~~~$206.963~(0.724)$~~~ & ~~~$216.201~(0.756)$~~~ & ~~~$218.305~(0.761)$~~~ \\\\ \\hline ~~~XCDM~~~ & ~~~$204.058~(0.719)$~~~ & ~~~$205.449~(0.721)$~~~ & ~~~$204.855~(0.719)$~~~ & ~~~$206.207~(0.721)$~~~ \\\\ \\hline ~~~CG~~~ & ~~~$204.050~(0.718)$~~~ & ~~~$205.365~(0.721)$~~~ & ~~~$204.552~(0.718)$~~~ & ~~~$205.930~(0.720)$~~~ \\\\ \\hline ~~~HDE~~~ & ~~~$204.070~(0.719)$~~~ & ~~~$205.496~(0.721)$~~~ & ~~~$204.228~(0.717)$~~~ & ~~~$205.614~(0.719)$~~~ \\\\ \\hline ~~~RDE~~~ & ~~~$204.058~(0.719)$~~~ & ~~~$205.449~(0.721)$~~~ & ~~~$205.875~(0.722)$~~~ & ~~~$207.498~(0.726)$~~~ \\\\ \\hline ~~~LP~~~ & ~~~$204.055~(0.721)$~~~ & ~~~$205.347~(0.723)$~~~ & ~~~$204.429~(0.720)$~~~ & ~~~$205.739~(0.722)$~~~ \\\\ \\hline ~~~CPL~~~ & ~~~$204.057~(0.721)$~~~ & ~~~$205.391~(0.723)$~~~ & ~~~$204.063~(0.719)$~~~ & ~~~$205.517~(0.721)$~~~ \\\\ \\hline ~~~GCG~~~ & ~~~$204.053~(0.721)$~~~ & ~~~$205.366~(0.723)$~~~ & ~~~$204.545~(0.720)$~~~ & ~~~$206.090~(0.723)$~~~ \\\\ \\hline\\hline \\end{tabular} \\end{center} \\end{table} \\begin{table} \\caption{The $\\chi_{min}^{2}$ and the $\\chi_{min}^{2}/dof$ (in the Parentheses) for the 10 models, where the Constitution dataset is used.} \\begin{center} \\label{table3} \\begin{tabular}{ccccc} \\hline\\hline ~~~Model~~~ & ~~~SNIa~~~ & ~~~SNIa+BAO~~~ & ~~~SNIa+CMB~~~ & ~~~SNIa+BAO+CMB~~~ \\\\ \\hline\\hline ~~~$\\Lambda$CDM~~~ & ~~~$465.604~(1.176)$~~~ & ~~~$466.902~(1.176)$~~~ & ~~~$466.316~(1.175)$~~~ & ~~~$467.525~(1.175)$~~~ \\\\ \\hline ~~~DGP~~~ & ~~~$466.122~(1.177)$~~~ & ~~~$467.433~(1.177)$~~~ & ~~~$498.264~(1.255)$~~~ & ~~~$500.368~(1.257)$~~~ \\\\ \\hline ~~~ADE~~~ & ~~~$466.275~(1.177)$~~~ & ~~~$467.580~(1.178)$~~~ & ~~~$483.675~(1.218)$~~~ & ~~~$485.585~(1.220)$~~~ \\\\ \\hline ~~~XCDM~~~ & ~~~$465.602~(1.179)$~~~ & ~~~$466.901~(1.179)$~~~ & ~~~$465.657~(1.176)$~~~ & ~~~$466.947~(1.176)$~~~ \\\\ \\hline ~~~CG~~~ & ~~~$465.293~(1.178)$~~~ & ~~~$466.564~(1.178)$~~~ & ~~~$465.591~(1.176)$~~~ & ~~~$466.891~(1.176)$~~~ \\\\ \\hline ~~~HDE~~~ & ~~~$465.769~(1.179)$~~~ & ~~~$467.080~(1.179)$~~~ & ~~~$466.030~(1.177)$~~~ & ~~~$467.384~(1.177)$~~~ \\\\ \\hline ~~~RDE~~~ & ~~~$465.602~(1.179)$~~~ & ~~~$466.901~(1.179)$~~~ & ~~~$472.841~(1.194)$~~~ & ~~~$474.466~(1.195)$~~~ \\\\ \\hline ~~~LP~~~ & ~~~$461.071~(1.170)$~~~ & ~~~$461.570~(1.169)$~~~ & ~~~$465.610~(1.179)$~~~ & ~~~$466.910~(1.179)$~~~ \\\\ \\hline ~~~CPL~~~ & ~~~$461.526~(1.171)$~~~ & ~~~$462.112~(1.170)$~~~ & ~~~$465.636~(1.179)$~~~ & ~~~$466.902~(1.179)$~~~ \\\\ \\hline ~~~GCG~~~ & ~~~$465.080~(1.180)$~~~ & ~~~$466.360~(1.181)$~~~ & ~~~$465.920~(1.180)$~~~ & ~~~$466.895~(1.179)$~~~ \\\\ \\hline\\hline \\end{tabular} \\end{center} \\end{table} \\begin{table} \\caption{The $\\chi_{min}^{2}$ and the $\\chi_{min}^{2}/dof$ (in the Parentheses) for the 10 models, where the ConstitutionT dataset is used.} \\begin{center} \\label{table4} \\begin{tabular}{ccccc} \\hline\\hline ~~~Model~~~ & ~~~SNIa~~~ & ~~~SNIa+BAO~~~ & ~~~SNIa+CMB~~~ & ~~~SNIa+BAO+CMB~~~ \\\\ \\hline\\hline ~~~$\\Lambda$CDM~~~ & ~~~$269.081~(0.743)$~~~ & ~~~$270.610~(0.745)$~~~ & ~~~$271.423~(0.748)$~~~ & ~~~$272.764~(0.749)$~~~ \\\\ \\hline ~~~DGP~~~ & ~~~$269.443~(0.744)$~~~ & ~~~$270.979~(0.747)$~~~ & ~~~$292.067~(0.805)$~~~ & ~~~$294.360~(0.809)$~~~ \\\\ \\hline ~~~ADE~~~ & ~~~$269.613~(0.745)$~~~ & ~~~$271.139~(0.747)$~~~ & ~~~$280.090~(0.772)$~~~ & ~~~$282.147~(0.775)$~~~ \\\\ \\hline ~~~XCDM~~~ & ~~~$269.066~(0.745)$~~~ & ~~~$270.599~(0.748)$~~~ & ~~~$269.245~(0.744)$~~~ & ~~~$270.754~(0.746)$~~~ \\\\ \\hline ~~~CG~~~ & ~~~$268.992~(0.745)$~~~ & ~~~$270.502~(0.747)$~~~ & ~~~$269.066~(0.743)$~~~ & ~~~$270.594~(0.745)$~~~ \\\\ \\hline ~~~HDE~~~ & ~~~$269.113~(0.745)$~~~ & ~~~$270.664~(0.748)$~~~ & ~~~$269.130~(0.743)$~~~ & ~~~$270.698~(0.746)$~~~ \\\\ \\hline ~~~RDE~~~ & ~~~$269.066~(0.745)$~~~ & ~~~$270.599~(0.748)$~~~ & ~~~$273.305~(0.755)$~~~ & ~~~$275.180~(0.758)$~~~ \\\\ \\hline ~~~LP~~~ & ~~~$268.812~(0.747)$~~~ & ~~~$270.021~(0.748)$~~~ & ~~~$269.068~(0.745)$~~~ & ~~~$270.603~(0.748)$~~~ \\\\ \\hline ~~~CPL~~~ & ~~~$268.900~(0.747)$~~~ & ~~~$270.147~(0.748)$~~~ & ~~~$269.122~(0.745)$~~~ & ~~~$270.676~(0.748)$~~~ \\\\ \\hline ~~~GCG~~~ & ~~~$268.980~(0.747)$~~~ & ~~~$270.473~(0.749)$~~~ & ~~~$269.069~(0.745)$~~~ & ~~~$270.599~(0.748)$~~~ \\\\ \\hline\\hline \\end{tabular} \\end{center} \\end{table} (1) For each DE model, the truncated SNIa datasets not only greatly reduce $\\chi _{min}^{2}$ and $\\chi _{min}^{2}/dof$, but also remove the tension between SNIa data and other cosmological observations. Since it is too prolix to describe the results of all 10 models, in the following we will just discuss the HDE model as an example. From table \\ref{table1} and table \\ref{table2}, we can see that the combined Union+BAO+CMB data gives a $\\chi_{min}^{2}=312.481$ and a $\\chi_{min}^{2}/dof=1.018$, while the combined UnionT+BAO+CMB data gives a $\\chi_{min}^{2}=205.614$ and a $\\chi_{min}^{2}/dof=0.719$. This means that by dropping the 21 outliers, The $\\chi_{min}^{2}$ and the $\\chi_{min}^{2}/dof$ of the HDE model can reduce 106.867 and 0.299, respectively. From table \\ref{table3} and table \\ref{table4}, similar results are obtained. The combined Constitution+BAO+CMB data gives a $\\chi_{min}^{2}=467.384$ and a $\\chi_{min}^{2}/dof=1.177$, while the combined ConstitutionT+BAO+CMB data gives a $\\chi_{min}^{2}=270.698$ and a $\\chi_{min}^{2}/dof=0.746$. This means that by dropping the 34 outliers, The $\\chi_{min}^{2}$ and the $\\chi_{min}^{2}/dof$ of the HDE model can reduce 196.686 and 0.431, respectively. It should be pointed out that for all these 10 models, the UnionT and the ConstitutionT dataset can greatly reduce the corresponding $\\chi _{min}^{2}$ and $\\chi _{min}^{2}/dof$. Moreover, the decreasing margin of $\\chi _{min}^{2}$ and $\\chi _{min}^{2}/dof$ for these 10 DE models are almost same. To further verify this conclusion, we also adopt the method of random truncation used in \\cite{ness07,wei10}. The method is very simple. First, we random select the outliers from the full Union dataset and the full Constitution dataset, respectively. Notice that the numbers of data points in these random ``UnionOut'' subsets are same as that in the original UnionOut samples, while the numbers of data points in these random ``ConstitutionOut'' subsets are same as that in the original ConstitutionOut samples. As in \\cite{ness07,wei10}, 500 random UnionOut subsets and 500 random ConstitutionOut subsets are selected. By subtracting these outliers from the Union dataset and the Constitution dataset, respectively, 500 random UnionT subsets and 500 random ConstitutionT subsets are obtained. Then, we perform best-fit analysis on the HDE model by using these random truncated SNIa subsets. Using the SNIa data only, we get the Mean of $\\chi_{min}^{2}$ for 500 random UnionOut subsets and 500 random ConstitutionOut subsets, respectively. Notice that the full Union set gives a $\\chi_{min}^{2}=310.827$ and the full Constitution set gives a $\\chi_{min}^{2}=465.769$, the differences between the $\\chi_{min}^{2}$ of the full SNIa dataset and the Mean of $\\chi_{min}^{2}$ of these random truncated SNIa subsets are also obtained. Next, we make a comparison for the original truncated SNIa subset and the random truncated SNIa subsets. The results are shown in table \\ref{table5}. It is found that the $\\chi_{min}^{2}$ of the HDE model can reduce 5 by dropping 1 data point in the original truncated SNIa subset, but can only reduce 1 by dropping 1 data point in the random truncated SNIa subset. Therefore, the original UnionOut and ConstitutionOut subsets are systematically different from the full Union and Constitution datasets, and the original truncated supernova datasets provide a significantly better model fitting than the full SNIa datasets. \\begin{table} \\caption{A comparison for the original truncated SNIa subset and the random truncated SNIa subsets. Notice that the full Union set gives a $\\chi_{min}^{2}=310.827$ and the full Constitution set gives a $\\chi_{min}^{2}=465.769$. Here $\\Delta N$ is the number of data points in outliers, $\\overline{{\\chi^2}}$ is the Mean of $\\chi_{min}^{2}$ for those random truncated SNIa subsets, and $\\Delta \\chi^2$ is the differences between the $\\chi_{min}^{2}$ of the full SNIa dataset and the $\\chi_{min}^{2}$ (or the Mean) of the random truncated SNIa subsets.} \\begin{center} \\label{table5} \\begin{tabular}{ccccc} \\hline\\hline ~~~The subset of outliers~~~ & ~~~$\\Delta N$~~~ & ~~~$\\chi_{min}^{2}$ or $\\overline{{\\chi^2}}$~~~ & ~~~$\\Delta \\chi^2$~~~ & ~~~$\\Delta \\chi^2/\\Delta N$~~~ \\\\ \\hline\\hline ~~~The original UnionOut subset~~~ & ~~~21~~~ & ~~~$204.070$~~~ & ~~~$106.757$~~~ & ~~~$5.084$~~~ \\\\ \\hline ~~~500 random UnionOut subsets~~~ & ~~~21 (Mean)~~~ & ~~~$289.617$ (Mean)~~~ & ~~~$21.210$ (Mean)~~~ & ~~~$1.010$ (Mean)~~~ \\\\ \\hline ~~~The original ConstitutionOut subset~~~ & ~~~34~~~ & ~~~$269.113$~~~ & ~~~$199.656$~~~ & ~~~$5.784$~~~ \\\\ \\hline ~~~500 random ConstitutionOut subsets~~~ & ~~~34 (Mean)~~~ & ~~~$425.763$ (Mean)~~~ & ~~~$40.006$ (Mean)~~~ & ~~~$1.177$ (Mean)~~~ \\\\ \\hline\\hline \\end{tabular} \\end{center} \\end{table} In \\cite{wei10}, by performing the best-fit analysis on the CPL model, Wei argued that the UnionT and the ConstitutionT datasets are fully consistent with the other cosmological observations. To verify this conclusion, more theoretical models should be taken into account. Here we study the HDE model as an example. In Fig.\\ref{fig1}, we plot the $1\\sigma$ and the $2\\sigma$ confidence level (CL) contours for the HDE model, where the Union and the UnionT datasets are used, respectively. From this figure, we find that the best-fit point for the Union data is outside the $2\\sigma$ confidence region given by the combined Union+BAO+CMB data, while the best-fit point for the UnionT data is inside the $1\\sigma$ confidence region given by the combined UnionT+BAO+CMB data. This means that the UnionT dataset is very useful to remove the tension between SNIa data and other cosmological observations. In addition, we also use the Constitution and the ConstitutionT datasets to plot the CL contours for the HDE model in Fig.\\ref{fig2}. It is found that the best-fit point for the Constitution data is outside the $2\\sigma$ confidence region given by the combined Constitution+BAO+CMB data, while the best-fit point for the ConstitutionT data is very close to the best-fit point for the combined ConstitutionT+BAO+CMB data. This means that the ConstitutionT dataset is also very helpful to remove the tension. Therefore, we conclude that the truncated SNIa datasets can remove the tension between SNIa data and other cosmological observations. \\begin{figure} \\includegraphics[scale=0.7, angle=0]{HDE_Union.eps} \\includegraphics[scale=0.7, angle=0]{HDE_Union_T.eps} \\caption{\\label{fig1} The $1\\sigma$ and the $2\\sigma$ CL contours for the HDE model. The left panel is plotted by using the Union dataset, while the right panel is plotted by using the UnionT dataset. For both these two panels, the blue dashed lines correspond to the constraints given by the SNIa data only, and the red solid lines correspond to the constraints given by the combined SNIa+BAO+CMB data. Moreover, we also plot the best-fit point for the SNIa data (blue point) and the best-fit point for the combined SNIa+BAO+CMB data (red star). It should be pointed out that, in the left panel, the best-fit point for the SNIa data is outside the $2\\sigma$ confidence region given by the combined SNIa+BAO+CMB data, while in the right panel, the best-fit point for the SNIa data is inside the $1\\sigma$ confidence region given by the combined SNIa+BAO+CMB data. This means that the UnionT dataset is very useful to remove the tension between SNIa data and other cosmological observations.} \\end{figure} \\begin{figure} \\includegraphics[scale=0.7, angle=0]{HDE_CFA.eps} \\includegraphics[scale=0.7, angle=0]{HDE_CFA_T.eps} \\caption{\\label{fig2} The same as in Fig.\\ref{fig1}, except for the cases of the Constitution and the ConstitutionT datasets. It should be mentioned that, in the left panel, the best-fit point for the SNIa only is outside the $2\\sigma$ confidence region given by the combined SNIa+BAO+CMB data, while in the right panel, the best-fit point for the SNIa only is very close to the best-fit point for the combined SNIa+BAO+CMB data. This means that the ConstitutionT dataset is very helpful to remove the tension between SNIa data and other cosmological observations.} \\end{figure} (2) The CMB data is very helpful to break the degeneracy among different parameters, and plays a very important role in distinguishing different DE models. As an example, we will compare the HDE model with the RDE model. By using the Constitution SNIa data alone, the CMB data alone, and the combined Constitution+CMB+BAO data, respectively, we plot the $1\\sigma$ and the $2\\sigma$ CL contours for the HDE and the RDE model in Fig.\\ref{fig3}. From this figure, we find that the shapes of CL contours given by the CMB data alone are quite different from that given by the SNIa data alone, and the CMB data is very helpful to break the degeneracy among different parameters. As shown in the left panel, the $1\\sigma$ CL contour of the HDE model given by the CMB data intersects to the $1\\sigma$ CL contour of the HDE model given by the SNIa data; while in the right panel, the $1\\sigma$ CL contour of the RDE model given by the CMB data does not intersect to the $1\\sigma$ CL contour of the RDE model given by the SNIa data. This fact explains why the HDE model performs much better in fitting the combined Constitution+BAO+CMB data than the RDE model. Besides, we also check the effect of the BAO data, and find that the $1\\sigma$ confidence region of the HDE model given by the BAO data alone can completely cover that given by the SNIa data. So the constraint given by the BAO data is quite weaker than that given by the SNIa and the BAO data. This conclusion can be further verified by using table \\ref{table3}. As seen in table \\ref{table3}, without adopting the CMB data, it is very difficult to distinguish the HDE Model from the RDE model. After adding the CMB data, it is seen that the $\\chi_{min}^{2}$ of the HDE Model given by the SNIa+CMB data is 6.811 smaller than the $\\chi_{min}^{2}$ of the RDE Model given by the SNIa+CMB data, while the $\\chi_{min}^{2}$ of the HDE Model given by the combined SNIa+CMB+BAO data is 7.082 smaller than the $\\chi_{min}^{2}$ of the RDE Model given by the combined SNIa+CMB+BAO data. \\begin{figure} \\includegraphics[scale=0.7, angle=0]{HDE_All_CFA.eps} \\includegraphics[scale=0.7, angle=0]{RDE_All_CFA.eps} \\caption{\\label{fig3} The $1\\sigma$ and the $2\\sigma$ CL contours for the HDE and the RDE model. The Constitution dataset is used to plot this figure. The left panel is plotted by using the HDE model, while the right panel is plotted by using the RDE model. For both these two panels, The blue dashed lines correspond to the constraints given by the SNIa data only, the black dotted lines correspond to the constraints given by the CMB data only, the red solid lines correspond to the constraints given by the combined SNIa+BAO+CMB data, and the red stars denote the best-fit point for the combined SNIa+BAO+CMB data. Since the shapes of CL contours given by the CMB data alone are quite different from that given by the SNIa data alone, the CMB data is very helpful to break the degeneracy among different parameters. As shown in the left panel, the $1\\sigma$ CL contour of the HDE model given by the CMB data intersects to the $1\\sigma$ CL contour of the HDE model given by the SNIa data; while in the right panel, the $1\\sigma$ CL contour of the RDE model given by the CMB data does not intersect to the $1\\sigma$ CL contour of the RDE model given by the SNIa data. This fact explains why the HDE model performs much better in fitting the combined Constitution+BAO+CMB data than the RDE model. Besides, we also check the effect of the BAO data, and find that the $1\\sigma$ confidence region of the HDE model given by the BAO data alone can completely cover that given by the SNIa data. For simplicity, the CL contours given by the BAO data are not plotted in this figure.} \\end{figure} In addition, by using the ConstitutionT SNIa data alone, the CMB data alone, and the combined ConstitutionT+CMB+BAO data, respectively, we also plot the $1\\sigma$ and the $2\\sigma$ CL contours for the HDE and the RDE model in Fig.\\ref{fig4}. Again, we see that the $1\\sigma$ CL contour of the HDE model given by the CMB data intersects to the $1\\sigma$ CL contour of the HDE model given by the SNIa data, while the $1\\sigma$ CL contour of the RDE model given by the CMB data does not intersect to the $1\\sigma$ CL contour of the RDE model given by the SNIa data. This fact explains why the HDE model performs much better in fitting the combined ConstitutionT+BAO+CMB data than the RDE model. Therefore, the CMB data is very helpful to break the degeneracy among different parameters, and plays a very important role in distinguishing different DE models. \\begin{figure} \\includegraphics[scale=0.7, angle=0]{HDE_All_CFA_T.eps} \\includegraphics[scale=0.7, angle=0]{RDE_All_CFA_T.eps} \\caption{\\label{fig4} The same as in Fig.\\ref{fig3}, except for the case of the ConstitutionT dataset. Notice that the shapes of CL contours given by the CMB data alone are quite different from that given by the SNIa data alone. As shown in the left panel, the $1\\sigma$ CL contour of the HDE model given by the CMB data intersects to the $1\\sigma$ CL contour of the HDE model given by the SNIa data; while in the right panel, the $1\\sigma$ CL contour of the RDE model given by the CMB data does not intersect to the $1\\sigma$ CL contour of the RDE model given by the SNIa data. This fact explains why the HDE model performs much better in fitting the combined ConstitutionT+BAO+CMB data than the RDE model.} \\end{figure} (3) The current observational data are still too limited to distinguish all DE models. First, we discuss three kinds of two-parameter models: the XCDM model, the CG model, and the HDE model. As seen in table \\ref{table3}, the $\\chi_{min}^{2}$ of these 3 models given by the combined Constitution+BAO+CMB data are 466.947, 466.891, and 467.384, respectively. That is to say, after taking into account the CMB data, the differences of the $\\chi_{min}^{2}$ of these 3 models are still smaller than 1. Notice that this result also holds true in table \\ref{table1}, table \\ref{table2}, and table \\ref{table4}. Therefore, one cannot judge which DE model is better. Then, we discuss three kinds of three-parameter models: the LP model, the CPL model, and the GCG model. As seen in table \\ref{table3}, the $\\chi_{min}^{2}$ of these 3 models given by the combined Constitution+BAO+CMB data are 466.910, 466.902, 466.895, respectively, and the differences of the $\\chi_{min}^{2}$ of these 3 models are even smaller than 0.1. Since this result also holds true in table \\ref{table1}, table \\ref{table2}, and table \\ref{table4}, it is also very difficult to distinguish these 3 models. Therefore, to distinguish DE models better, more high-quality observational data are needed." }, "0910/0910.2712_arXiv.txt": { "abstract": "We combine a cosmological reionization simulation with box size of 100$\\hMpc$ on a side and a Monte Carlo \\lya radiative transfer code to model Lyman Alpha Emitters (LAEs) at $z\\sim$5.7. The model introduces \\lya radiative transfer as the single factor for transforming the intrinsic \\lya emission properties into the observed ones. Spatial diffusion of \\lya photons from radiative transfer results in extended \\lya emission and only the central part with high surface brightness can be observed. Because of radiative transfer, the appearance of LAEs depends on density and velocity structures in circumgalactic and intergalactic media as well as the viewing angle, which leads to a broad distribution of apparent (observed) \\lya luminosity for a given intrinsic \\lya luminosity. Radiative transfer also causes frequency diffusion of \\lya photons. The resultant \\lya line is asymmetric with a red tail. The peak of the \\lya line shifts towards longer wavelength and the shift is anti-correlated with the apparent to intrinsic \\lya luminosity ratio. The simple radiative transfer model provides a new framework for studying LAEs. It is able to explain an array of observed properties of $z\\sim$5.7 LAEs in Ouchi et al. (2008), producing \\lya spectra, morphology, and apparent \\lya luminosity function (LF) similar to those seen in observation. The broad distribution of apparent \\lya luminosity at fixed UV luminosity provides a natural explanation for the observed UV LF, especially the turnover towards the low luminosity end. The model also reproduces the observed distribution of \\lya equivalent width (EW) and explains the deficit of UV bright, high EW sources. Because of the broad distribution of the apparent to intrinsic \\lya luminosity ratio, the model predicts effective duty cycles and \\lya escape fractions for LAEs. ", "introduction": "More than four decades ago, \\citet{Partridge67} proposed that prominent \\lya emission reprocessed from ionizing photons of young stars in galaxies can be used to detect high-redshift galaxies. The first successful detections of high-redshift \\lya emitting galaxies, or \\lya emitters (LAEs), were made $\\sim$ 30 years later \\citep[e.g.,][]{Hu96,Cowie98,Dey98,Hu98,Hu99}. Recently, important advances have been made on the observational front to detect LAEs at $z\\gtrsim 6$ \\citep[e.g.,][]{Hu98,Hu02,Hu04,Hu05,Hu06,Rhoads03,Malhotra04,Horton04,Stern05, Kashikawa06,Shimasaku06,Iye06,Cuby07,Ouchi07,Ouchi08,Stark07a,Nilsson07, Willis08,Ota08}. LAEs can be efficiently detected through narrow-band imaging or with integral-field-units (IFU) spectroscopy. Owing to the high efficiency of target detection, LAEs naturally become objects for large surveys of high-redshift galaxies. Besides providing clues to the formation and evolution of galaxies at the time when the universe was still young, LAEs are an important tracer of the large-scale structure. The clustering of LAEs may be used to constrain cosmological parameters. In particular, the large-volume surveys such as the Hobby-Eberly Telescope Dark Energy Experiment (HETDEX; \\citealt{Hill08}) will enable the detection of the baryon acoustic oscillations (BAO) features \\citep[e.g.,][]{Eisenstein05} in the LAE power spectrum. The BAO and the shape of the power spectrum can be used to measure the expansion history of the universe at early epochs ($z\\sim 3$), which constrains the evolution of dark energy and the curvature of the universe. LAEs are also a key probe of the high-redshift intergalactic medium (IGM), especially across the reionization epoch. The use of LAEs to learn about reionization has been the subject of intense study \\citep[e.g.,][]{Rees98,Miralda98,Haiman99,Santos04,Haiman05,Dijkstra07a, Wyithe07}. Suitably devised statistics, including luminosity function (LF) and correlation functions of LAEs, can be used to constrain the neutral fraction of the IGM during reionization \\citep[]{Malhotra04,Haiman05,Kashikawa06,Furlanetto06,Dijkstra07b,McQuinn07, Mesinger08,Iliev08,Dayal08,Dayal09}. By comparing the LFs of $z\\sim 5.7$ and $z\\sim 6.5$ LAEs, \\citet{Malhotra04} conclude that reionization was largely complete at $z\\sim 6.5$ (also see \\citealt{Dijkstra07b}). \\citet{McQuinn07} show that, with the angular correlation function of the 58 available $z\\sim 6.6$ LAEs in the Subaru Deep Field \\citep{Kashikawa06}, limits may be placed on the IGM neutral fraction, favoring a fully ionized universe at $z\\sim 6.6$. However, none of the previous work of LAEs mentioned above used reionization simulations with concurrent treatment of hydrodynamics plus radiative transfer of ionizing photons and \\lya photons. Hydrodynamic and radiative transfer simulations provide realistic neutral gas distributions, and \\lya radiative transfer yields detailed properties of the \\lya emission. Realistic \\lya radiative transfer calculations have been applied to high-redshift LAEs in cosmological simulations \\citep[e.g.,][]{Tasitsiomi06}. The application, however, is limited to a few individual sources, which do not form a sample for statistical study. \\citet{McQuinn07} and \\citet{Iliev08} studied a sample of LAEs in reionization simulations with cosmological volume. However, the radiative transfer of \\lya photons is treated in a simplistic way in their study: the observed \\lya spectrum is modeled as the intrinsic line profile modified by $\\exp(-\\tau_\\nu)$, where $\\tau_\\nu$ is the optical depth at frequency $\\nu$ along the line of sight. Although this $\\exp(-\\tau_\\nu)$ model can yield insights into the properties of the observed \\lya emission, such as the effect of IGM on the observability of LAEs, it is far from a complete description of the radiative transfer of \\lya photons. First, during the propagation, \\lya photons experience frequency diffusion, which is neglected by the simple $\\exp(-\\tau_\\nu)$ model. The $\\exp(-\\tau_\\nu)$ model removes \\lya photons at a given frequency according to the \\lya optical depth, and no frequency change occurs for any \\lya photon, therefore it does not yield correct \\lya spectra. Second, the simple $\\exp(-\\tau_\\nu)$ model does not account for the spatial diffusion of \\lya photons either. LAEs in this model appear as point sources in \\lya and there is no surface brightness information. Even if \\lya photons start from a point source, spatial diffusion due to radiative transfer would lead to an extended source. Observationally, LAEs indeed appear to be extended and they are defined by a surface brightness threshold in the narrow-band image \\citep[e.g.,][]{Ouchi08}. Therefore, although the simple $\\exp(-\\tau_\\nu)$ model may provide useful insight, it likely falls short for predicting the detailed properties of the observed \\lya emission from LAEs. To correctly understand high-redshift LAEs and use them for cosmological study, a full calculation of radiative transfer of \\lya photons for a large sample of LAEs in cosmological reionization simulation is necessary, as will be evident later. In this work, we aim to perform detailed radiative transfer calculation of \\lya photons \\citep{Zheng02} from LAEs in a self-consistent fashion through radiation-hydrodynamic reionization simulations \\citep{Trac08}. For this paper, we focus on studying statistical properties of $z\\sim 5.7$ LAEs and show how the radiative transfer calculation aids our understanding of the observed properties of LAEs. The clustering properties of LAEs from this study will be presented in another paper (Paper II; Zheng et al. in prep.). The paper is organized as follows. In \\S~2 we review the cosmological reionization simulation used in this work and in \\S~3 we describe the \\lya radiative transfer calculation. In \\S~4, we study in details the \\lya emission from an individual source chosen from the simulation box to gain a general view of the effect of \\lya radiative transfer on the appearance of LAEs. Then, we present the statistical properties of LAEs in \\S~5, including their spectra and luminosity, from our modeling of an ensemble of sources in the simulation box. We compare our modeling results with observations for $z\\sim 5.7$ LAEs and discuss the implications in our understanding of LAEs. \\S~7 is devoted to identifying important physical factors in shaping the observed \\lya emission of LAEs. We summarize and discuss the results in \\S~8. ", "conclusions": "\\subsection{Summary of Main Results} We perform a full \\lya radiative transfer calculation with a Monte Carlo code \\citep{Zheng02} to study LAEs in a cosmological volume. The LAE sources and the physical properties of neutral hydrogen gas are taken from the $z\\sim 5.7$ outputs of a cosmological reionization simulation \\citep{Trac08}, which solves the coupled evolution of the dark matter, baryons, and ionizing radiation in a box of 100$\\hMpc$ (comoving) on a side. The large volume of the simulation allows a statistical study of $z\\sim 5.7$ LAEs. Radiative transfer of \\lya photons in the IGM environment around LAEs, which leads to both frequency and spatial diffusion of \\lya photons, turns out to play a crucial role in determining the observability of LAEs and in understanding the observed properties of LAEs. Although the radiative transfer calculation is computationally costly, the LAE model we present in this paper is rather simple. The UV or intrinsic \\lya luminosity is assumed to be proportional to the SFR, which is tightly coupled to halo mass in the reionization simulation we use. That is, we essentially adopt a constant mass to light ratio, where mass is halo mass and light is either UV or \\lya. All we do is to add the physics of \\lya radiative transfer into the model to obtain the observed properties of LAEs. That is, we introduce the radiative transfer of \\lya photons in the IGM as the single factor responsible for transforming the intrinsic \\lya emission properties to the observed ones. Our model produces IFU-like data cube that covers the extent of the simulation box, which allows mock observations to be made. With the \\lya image contracted from this data cube, we follow typical observational procedures \\citep[e.g.,][]{Ouchi08} to identify LAEs and then extract their \\lya spectra. Initially \\lya photons are produced inside the star formation region. Therefore the intrinsic \\lya sources are expected to be similar in size as the UV sources, which are compact ($\\lesssim 1$ kpc; \\citealt{Taniguchi09}). We find that an intrinsically point-like \\lya source becomes extended as a consequence of resonant scatterings of \\lya photons (spatial diffusion). The scatterings of \\lya photons do not destroy them and all \\lya photons escape in the end. However, observationally, only the central part of the extended source can be detected as a consequence of the limit set by the surface brightness threshold. The scatterings of \\lya photons also cause the frequency of \\lya photons to change (frequency diffusion). The resultant \\lya spectra from the central aperture do not have a simple relation to the initial profile, which is assumed to be Gaussian in our model. Our results from full radiative transfer calculations show a clear difference from a simple treatment of \\lya radiative transfer, namely the $\\exp(-\\tau_\\nu)$ model, widely adopted in previous work, which modifies the intrinsic \\lya spectrum by multiplying the line-of-sight transmission determined by the optical depth at each frequency. The observed \\lya spectrum of an LAE in our model shows a clear asymmetry, skewed towards red. Although the $\\exp(-\\tau_\\nu)$ model produces the same qualitative feature, the predicted line profile, the frequency shift, and the total flux are all significantly different from our results. While the spectrum of the $\\exp(-\\tau_\\nu)$ model we present is essentially the intrinsic one truncated below a certain wavelength (but see Figures 14 and 15 in \\citealt{Iliev08} for more complex line shapes, probably caused by different assumptions in the $\\exp(-\\tau_\\nu)$ model), the observed \\lya spectrum in our model can have contributions from photons with frequency much redder than initial photons, a result of the scattering-caused frequency diffusion. We find that the redward shift of the Lya line induced by radiative transfer is usually a few times the intrinsic line width, with a distribution that peaks at about three times. The asymmetry and shift of the \\lya line do not indicate the presence of any winds, but they arise from the structure of the halo infall and Hubble expansion around the sources. If one were to infer the wind velocity, if there is any, from comparing the relative shift in the \\lya line and an optically-thin line, one has to keep in mind the \\lya radiative transfer effect. For example, the observationally inferred velocity of the receding winds would be overestimated by $\\sim 100\\kms$ or more if the effect is not taken into account. As a consequence of the frequency diffusion and spatial diffusion, our model predicts a much higher observed \\lya flux than the $\\exp(-\\tau_\\nu)$ model. At a fixed intrinsic \\lya luminosity (i.e., fixed host halo mass), the observed (apparent) luminosity is broadly distributed. The shift in the \\lya line peak and the ratio of the apparent to intrinsic \\lya luminosity appear to be anti-correlated. The distributions of the line peak shift and the ratio of apparent to intrinsic \\lya luminosity, and their correlation, all result from the dependence of the \\lya radiative transfer on the IGM environments around sources. It would be interesting and extremely useful if we could make use of the full information in the observed \\lya properties to infer the intrinsic ones, and we reserve such an investigation for future work. Although our model predicts a much higher observed \\lya flux than the $\\exp(-\\tau_\\nu)$ model, it still leads to a highly suppressed \\lya flux, compared with the intrinsic one. The suppression factor depends on the assumed line width of the intrinsic \\lya spectra. For the line width assumed in our model (given by halo virial temperature), we find that, with respect to the intrinsic \\lya LF of LAEs, the observed (apparent) \\lya LF shift towards the low luminosity end by roughly one order of magnitude in luminosity. For comparison, the $\\exp(-\\tau_\\nu)$ model would shift by two orders of magnitude in luminosity. We make comparisons between the $z\\sim 5.7$ LAEs in our model and those observed in SXDS \\citep{Ouchi08}. The sizes, morphologies, \\lya line profiles of the model LAEs are remarkably similar to the observed ones. For the \\lya LF, UV LF, and \\lya EW distribution, our model can successfully reproduce the observations and provide physical explanations for various observed features. After an overall adjustment of a factor of $\\sim 5$ in luminosity, the \\lya LF of model LAEs matches well with observation. The adjustment reflects our incomplete knowledge in the stellar IMF at high redshift, the uncertainty in the model SFR, and the lack of information on the intrinsic \\lya line profile. According to our model, there is no one-to-one map between the intrinsic and the observed \\lya luminosity. In other words, there is a large scatter in the relation between the apparent and intrinsic luminosities. At a fixed observed luminosity, LAEs can differ by one order of magnitude in the intrinsic luminosity (Fig.~\\ref{fig:LL}$c$). This large scatter has to be taken into account when interpreting the observed \\lya LF and linking the observed LAEs to their host halos. For the UV LF of observed LAEs, our model prediction shows a good agreement with observation. In particular, the turnover of the UV LF towards the low luminosity end seen in high-$z$ ($z\\sim$ 3--6) LAEs are well reproduced. The key to interpret the shape of the UV LF is that observed LAEs are sources with observed (apparent) \\lya luminosity above certain threshold. The turnover reflects that for LAEs with low UV luminosity (or low intrinsic \\lya luminosity, or low halo mass), the probability for the observed \\lya luminosity to exceed the observation threshold is low, a consequence of the broad distribution of apparent \\lya luminosity at a given intrinsic \\lya luminosity. The full UV LF for sources in our model (i.e., without imposing the observation \\lya luminosity threshold) agrees well with the nonduplicated sum of the observed UV LFs of LAEs and $i$--dropout galaxies at $z\\sim 6$. The observed distribution of \\lya EW as a function of UV luminosity is also reproduced in the model. We note that in our model all the sources have the same {\\it intrinsic} EW and the distribution of the {\\it observed} values of EW at fixed UV luminosity is purely caused by the environment-dependent radiative transfer effect. At a fixed UV luminosity, the distribution of observed (apparent) EW is a decreasing function toward high values. The observational trend of lacking UV bright, high EW sources \\citep[e.g.,][]{Ando06} is naturally explained by our model in that such sources lie in a low probability corner --- a combination of the drop of the UV LF toward high luminosity and the drop of the apparent EW distribution function toward high EW value. LAE surveys with large volume will test the interpretation. Therefore, the observed properties of LAEs can be explained by simply invoking \\lya radiative transfer: the effects of the local IGM environment, depending mainly on the gas density and line-of-sight velocity and their line-of-sight gradients, lead to the distribution of observed \\lya emission properties at fixed intrinsic \\lya luminosity. This environmental selection also causes new features in the clustering of LAEs that we will study in Paper II. \\subsection{Implications and Discussion} Our interpretation of the observations of LAEs does not invoke any mass dependent dust absorption, which is in contrast to many previous models \\citep[e.g.,][]{Dayal10}. Uniformly distributed dust efficiently absorbs \\lya photons, since the large number of resonant scatterings increase the path length. There is much less attenuation when the dust is in gas clumps and \\lya photons bounce off the cloud surfaces \\citep{Neufeld91,Hansen06}. Optical, UV, and \\lya observations of local star-forming galaxies provide evidence that ISM kinematics and geometry play a more significant role than dust in affecting the \\lya emission \\citep[e.g.,][]{Giavalisco96,Keel05,Atek08, Atek09}. Our model successfully reproduces the observed UV LF of LAEs by incorporating only a mass independent effective UV extinction of at most 0.3 mag. We conclude that any mass dependent dust effects are not likely to play a substantial role to determine the observed properties of LAEs, compared to \\lya radiative transfer effects. Our model also has important implications for the duty cycle and the \\lya escape fraction of LAEs. The theoretically predicted (intrinsic) \\lya LF, which essentially is the halo mass function, is substantially higher than the observed one. Two scenarios have been introduced to address this problem, the duty cycle and the \\lya escape fraction scenarios \\citep[e.g.,][]{Stark07b,Nagamine08}. In the duty cycle scenario, LAEs are short-lived and a fraction of all galaxies are active as LAEs at any given time, lowering the amplitude of the \\lya LF. In the \\lya escape fraction scenario, only a fixed fraction of \\lya photons escape from the source and the overprediction problem is solved by shifting the LF towards the low luminosity end. To conserve the number density of LAEs of a given sample, the masses of host halos in the duty cycle scenario would be on average lower than those in the escape fraction scenario. As a consequence, the clustering of LAEs would be different in the two scenarios, with a stronger clustering in the escape fraction scenario. \\citet{Nagamine08} find that LAE clustering measurements from observations are in favor of their duty cycle scenario. In our model, \\lya photons all escape after a large number of scatterings. The \\lya escape fraction, in its literal meaning, is therefore unity. However, only the central part of the extended \\lya emission of LAEs can be observed, which gives rise to an {\\it apparent} or {\\it effective} \\lya escape fraction. Since the observed \\lya luminosity has a broad distribution at a fixed intrinsic \\lya luminosity, our model predicts a broad distribution of the effective \\lya escape fraction rather than a single value. In our model, no duty cycle parameter is introduced. Since halos of the same mass have similar SFR in our model, the corresponding intrinsic \\lya luminosities are the same, i.e., \\lya emission does not come from a fraction of halos. However, an {\\it apparent} or {\\it effective} duty cycle arises as a result of the selection effect caused by \\lya radiative transfer (a broad distribution of observed \\lya luminosity at a fixed intrinsic \\lya luminosity) and a \\lya luminosity threshold in observation. This can be seen from comparing the UV LF for all galaxies (dropout galaxies and LAEs) and that for LAEs (Fig.~\\ref{fig:uvLF}), which can be described as that at a fixed UV luminosity (or halo mass) only a fraction of all the galaxies are observed as LAEs. This effective duty cycle does not have the physical meaning in its original form. Moreover, it is not a constant, since it changes with UV luminosity (Fig.~\\ref{fig:uvLF}). \\citet{Tilvi09} present an LAE model in which \\lya luminosity (or SFR) is related to the halo mass accretion rate, rather than halo mass, and the model naturally gives rise to the duty cycle of LAEs. The duty cycle in their model, however, has its original meaning, in direct contrast with our model. Our model still ties the intrinsic \\lya luminosity (SFR) to halo mass and let the \\lya radiative transfer do the work of converting it to observed \\lya luminosity. Because of the large scatter between the observed \\lya luminosity and the intrinsic one (or halo mass) in our model, for a sample of LAEs above a \\lya luminosity threshold, some of them can reside in halos with mass smaller than the threshold mass above which halos have the same number density as LAEs. So the effective duty cycle in our model has the effect of lowering the clustering amplitude of LAEs. The main uncertainties in our model are the stellar IMF, the SFR, and the intrinsic \\lya line profile. The first two are general uncertainties for any model. The IMF at high-$z$ is neither well constrained observationally nor well understood theoretically. The SFR in galaxy formation model is related to the complex gas physics that we do not have a satisfactory understanding. Changing IMF or SFR would change the details in the reionization process and therefore change the gas properties (e.g., neutral fraction and temperature distribution) at a fixed redshift. For the reionization history itself, the escape fraction of ionizing photons adds a further uncertainty. Even though we focus on LAEs at $z\\sim 5.7$, when reionization is almost complete, different reionization histories can still leave different imprints on the gas distribution. For example, the IGM temperature in a region is correlated to the time when this region is reionized and heated \\citep[e.g.,][]{Trac08}. A detailed study is needed to investigate the effect of inhomogeneous IGM temperature distribution on \\lya radiative transfer. To be fully self-consistent, for any change in the IMF, SFR, and escape fraction of ionizing photons, one has to re-run the reionization simulation to solve the density, velocity, and temperature distributions of neutral gas and then perform the \\lya radiative transfer calculation. If the IMF, SFR, and escape fraction of ionizing photons change in a way to maintain the same reionization history, the effect of the IMF and SFR change can be largely characterized by an overall scaling in UV or intrinsic \\lya luminosity and one does not need to redo the \\lya radiative transfer calculation. For simplicity and to avoid extensive computations for reionization and radiative transfer simulation, we adopt such a scenario in this paper. Changing the width of the intrinsic \\lya line profile leads to changes in the distribution of the apparent to intrinsic \\lya luminosity ratio. Although we cast the effect as an overall scaling in the apparent \\lya luminosity for the \\lya LF, the intrinsic \\lya line profile is important in many aspects of the \\lya observation (image, spectra, etc) and its effect deserves a detailed investigation. \\lya radiative transfer calculation with high resolution hydrodynamic simulations for individual LAEs are necessary to shed light on the intrinsic \\lya line profile \\citep[e.g.,][]{Laursen07}. Changing the IMF affects the flux of ionizing photons more than that of the $\\sim$ 1500\\AA UV photons, which leads to a change in the ratio of the intrinsic \\lya luminosity to UV luminosity, or the intrinsic \\lya EW. Intrinsic \\lya line profile and dust play a role in converting the intrinsic EW to apparent EW, with the former affecting the \\lya luminosity and the latter adding extinction to UV luminosity. Therefore, the full distribution of observed \\lya EW and UV luminosity of LAEs can give constraints on the IMF, SFR, dust, and intrinsic \\lya profile. Although we have identified key environment factors in shaping the observational properties of LAEs, the dependence of radiative transfer on environments deserves a further study to understand the details of \\lya scatterings in the surrounding regions of LAEs. Our \\lya radiative transfer calculation relies on the gas distribution and properties from the cosmological reionization simulation. The radiative transfer of ionizing photons with a ray tracing algorithm is crucial in determining the state of gas. We have tested our LAE model for a reionization simulation with improved ray tracing algorithm (Trac et al. in prep.) in a small box (25$\\hMpc$ on a side). We find that the results presented in this paper are robust. \\lya radiative transfer through the surrounding circumgalactic and intergalactic media is a physical process that likely plays an important role in galaxies at all redshifts. It has to be taken into account for modeling LAEs and for interpreting observations. Our model is rather {\\it simple} and can naturally explain an array of observations of LAEs, which make it extremely attractive. It is interesting to see how well it does in interpreting observations of LAEs at lower redshifts (e.g., $z\\sim 3$). We also plan to apply it to the era of the late stage of reionization to study how to use LAEs to constrain reionization." }, "0910/0910.2238_arXiv.txt": { "abstract": "We investigate how the removal of interstellar material by stellar feedback limits the efficiency of star formation in molecular clouds and how this determines the shape of the mass function of young star clusters. In particular, we derive relations between the power-law exponents of the mass functions of the clouds and clusters in the limiting regimes in which the feedback is energy-driven and momentum-driven, corresponding to minimum and maximum radiative losses, and likely to bracket all realistic cases. We find good agreement between the predicted and observed exponents, especially for momentum-driven feedback, provided the protoclusters have roughly constant mean surface density, as indicated by observations of the star-forming clumps within molecular clouds. We also consider a variety of specific feedback mechanisms, concluding that \\hii\\ regions inflated by radiation pressure predominate in massive protoclusters, a momentum-limited process when photons can escape after only a few interactions with dust grains. We show in this case that the star formation efficiency depends on the masses and sizes of the protoclusters only through their mean surface density, thus ensuring consistency between the observed exponents of the mass functions of the clouds and clusters. Our numerical estimate of this efficiency is also consistent with observations. ", "introduction": "Most stars form in protoclusters in dense molecular clumps (\\citealt{lada03}; \\citealt{mckee07b}). The energy and momentum injected by young stars then removes the remaining interstellar material (ISM), thus ending further star formation and reducing the gravitational binding energy of the protoclusters. This feedback limits the efficiency of star formation---the ratio of final stellar mass to initial interstellar mass---to only $20-30\\%$, and leaves many protoclusters unbound, with their constituent stars free to disperse. Even those protoclusters that survive will lose some stars by ISM removal and subsequent processes. Two of the best probes of these formation and disruption processes are the mass functions of molecular clouds and young star clusters, defined as the number of objects per unit mass, $\\psi(M) \\equiv dN/dM$. For molecular clouds, the best-studied galaxies are the Milky Way and the Large Magellanic Cloud (LMC), while for star clusters, they are the Antennae and the LMC. In these and other cases, the observed mass functions can be represented by power laws, $\\psi(M) \\propto M^{\\beta}$, from $10^4M_{\\odot}$ or below to $10^6M_{\\odot}$ or above. Giant molecular clouds (GMCs) identified in CO surveys have $\\beta \\approx -1.7$ \\citep{rosolowsky05b, blitz07a, fukui08a}. This exponent is also found for massive self-gravitating clumps within GMCs, the formation sites of star clusters, whether they are identified by CO emission \\citep{bertoldi92} or higher-density tracers such as C$^{18}$O, $^{13}$CO, and thermal dust emission \\citep{reid06b, munoz07a, wong08a}. Young star clusters have $\\beta \\approx -2.0$ \\citep{elmegreen97a, mckee97, zhang99b, dowell08a, fall09a, chandar09a}. The similar exponents for clouds and clusters indicate that the efficiency of star formation and probability of disruption are at most weak functions of mass. This conclusion is reinforced by the fact that $\\beta$ is the same for $10^7-10^8$ yr-old clusters as it is for $10^6-10^7$ yr-old clusters \\citep{zhang99b, fall09a, chandar09a}. These empirical results may at first seem puzzling. Low-mass protoclusters have lower binding energy per unit mass and should therefore be easier to disrupt than high-mass protoclusters. Indeed, several authors have proposed that feedback would cause a bend in the mass function of young clusters at $M \\sim 10^5M_{\\odot}$, motivated in part by the well-known turnover in the mass function of old globular clusters \\citep{kroupa02a, baumgardt08a, parmentier08a}. For young clusters, such a feature is not observed (as noted above), while for globular clusters, it arises from almost any initial conditions as a consequence of stellar escape driven by two-body relaxation over $\\sim 10^{10}$~yr \\citep[and references therein]{fall01a, mclaughlin08a}. Nevertheless, we are left with an important question: What are the physical reasons for the observed similarity of the mass functions of molecular clouds and young star clusters? The goal of this Letter is to answer this question. In Section~\\ref{energymomentum}, we derive some general relations between the mass functions of clouds and clusters. In Section~\\ref{sec:efficiency}, we review a variety of specific feedback processes and estimate the star formation efficiency for radiation pressure, the dominant process in massive, compact protoclusters. We summarize in Section~\\ref{sec:conclusion}. ", "conclusions": "\\label{sec:conclusion} This Letter contains two main results. The first is the relation between the power-law exponents of the mass functions of molecular clouds and young star clusters, $\\beta_o$ and $\\beta_*$, in the limiting regimes in which stellar feedback is energy-driven and momentum-driven, Equations (\\ref{betastar:energy}) and (\\ref{betastar:momentum}), which bracket all realistic cases. The predicted $\\beta_*$ depends significantly on the initial size-mass relation of the protoclusters. We find good agreement between the predicted and observed $\\beta_*$, especially for momentum-driven feedback, for $\\Sigma \\propto M/R_h^2 \\approx {\\rm constant}$, the relation indicated by observations of gas-dominated protoclusters. In this case, the star formation efficiency is independent of protocluster mass, ensuring that the fraction of clusters that remain gravitationally bound following ISM removal is also independent of mass. The second main result is an estimate of the star formation efficiency in protoclusters regulated by radiation pressure, Equations (\\ref{efficiency}) and (\\ref{Sigma_crit}). This is likely to be the dominant feedback process in massive protoclusters. We show that ${\\cal E}$ depends on $M$ and $R_h$ only through the mean surface density $\\Sigma$, which in turn guarantees consistency between the observed power-law exponents of the mass functions of molecular clouds and young star clusters according to our general relations. For $\\Sigma \\sim 1$~g~cm$^{-2}$, we estimate ${\\cal E} \\sim 0.3$, in satisfactory agreement with observations." }, "0910/0910.5418_arXiv.txt": { "abstract": "{This paper reports the results obtained on the photometric redshifts measurement and accuracy, and cluster tomography in the ESO Distant Cluster Survey (EDisCS) fields.} {We present the methods used to determine photometric redshifts to discriminate between member and non-member galaxies and reduce the contamination by faint stars in subsequent spectroscopic studies.} {Photometric redshifts were computed using two independent codes both based on standard spectral energy distribution (SED) fitting methods ($Hyperz$ and G. Rudnick's code). Simulations were used to determine the redshift regions for which a reliable determination of photometric redshifts was expected. The accuracy of the photometric redshifts was assessed by comparing our estimates with the spectroscopic redshifts of $\\sim 1400$ galaxies in the $0.3\\le z\\le 1.0$ domain. The accuracy expected for galaxies fainter than the spectroscopic control sample was estimated using a degraded version of the photometric catalog for the spectroscopic sample. } {The accuracy of photometric redshifts is typically $\\sigma(\\Delta z/(1+z))\\sim 0.05 \\pm 0.01$, depending on the field, the filter set, and the spectral type of the galaxies. The quality of the photometric redshifts degrades by a factor of two in $\\sigma(\\Delta z/(1+z))$ between the brightest ($I\\ltapprox$22) and the faintest ($I\\sim$24-24.5) galaxies in the EDisCS sample. The photometric determination of cluster redshifts in the EDisCS fields using a simple algorithm based on \\zphot is in excellent agreement with the spectroscopic values, such that $\\delta z \\sim$0.03-0.04 in the high-z sample and $\\delta z \\sim$0.05 in the low-z sample, i.e. the \\zphot cluster redshifts are at least a factor $\\sim(1+z)$ more accurate than the measurements of \\zphot for individual galaxies. We also developed a method that uses both photometric redshift codes jointly to reject interlopers at magnitudes fainter than the spectroscopic limit. When applied to the spectroscopic sample, this method rejects $\\sim 50-90\\%$ of all spectroscopically confirmed non-members, while retaining $\\gtrsim 90\\%$ of all confirmed members.} {Photometric redshifts are found to be particularly useful for the identification and study of clusters of galaxies in large surveys. They enable efficient and complete pre-selection of cluster members for spectroscopy, allow accurate determinations of the cluster redshifts based on photometry alone, and provide a means of determining cluster membership, especially for bright sources. } ", "introduction": " ", "conclusions": "Discussion and Conclusions} We have used two independent codes to compute photometric redshifts: {\\it Hyperz} and GR code. In general, the two codes yield rather similar results, either on a cluster-by-cluster basis or as a function of the filter set and spectral type, of typically $\\sigma(\\Delta z/(1+z))\\sim 0.05$ to 0.06. {\\it Hyperz} results are found to be slightly more accurate than GR's ones in general, by $\\ltapprox$20\\% in $\\sigma(\\Delta z/(1+z))$. The quality achieved by both codes is consistent with the expectations derived from ``ideal'' simulations. An interesting trend is that the quality of both codes is highly correlated, in the sense that the highest and lowest quality results, in terms of $\\sigma(\\Delta z/(1+z))$, and systematics are found for the same clusters. This trend cannot be due to the use of an incomplete or imperfect template set, as suggested by other authors (Ilbert et al.\\ 2006), because in such a case, we should expect the same systematic behavior in all fields, given a filter set, as discussed in \\S~\\ref{simus}. In contrast, different systematics are observed in the different fields, which are found to be almost equal for the two independent \\zphot codes. This behavior suggests that the origin of the systematic errors is more likely to be associated with small residuals in the input photometry rather than the \\zphot templates and codes. Indeed, small zero-point shifts of $\\ltapprox$0.05 magnitudes cannot be excluded, in particular for the near-IR data. Photometric redshifts are found to be particularly useful in the identification and study of galaxy clusters in large surveys. The determination of cluster redshifts in the EDisCS fields using a simple algorithm based on \\zphot is highly accurate. Indeed, the differences between photometric and spectroscopic values are found to be small, typically ranging between $\\delta z \\sim$0.03-0.04 in the high-z sample and $\\delta z \\sim$0.05 in the low-z sample. This is at least a factor $\\sim(1+z)$ more accurate than the determination of \\zphot for individual galaxies. The accuracy is more sensitive to the filter set used rather than the redshift of the cluster. The systematic lower quality results for the low-z sample was somewhat expected from the simulations presented in \\S~\\ref{simus}. Tomography based on \\zphot could be used in searches for clusters along the line-of-sight, using redshift steps optimized to be close in value to the typical difference between photometric and spectroscopic z$_{cluster}$ to maximize the contrast between members and non-member galaxies (in this case, $\\Delta z \\sim 0.05$). The cluster membership criterion presented in Sect.~\\ref{member} has been used to extend the spectroscopic studies of cluster galaxies to fainter limits in magnitude (e.g. De Lucia et al.\\ 2004, White et al.\\ 2005, Clowe et al.\\ 2006, Poggianti et al.\\ 2006, De Lucia et al.\\ 2007, Desai et al.\\ 2007, Rudnick et al.\\ 2009). In conclusion, photometric redshifts are useful tools for studying galaxy clusters. They enable efficient and complete pre-selection of cluster members for spectroscopy, allow accurate determinations of the cluster redshifts based on photometry alone, provide a means of determining cluster membership, especially for bright sources, and can be used to search for galaxy clusters." }, "0910/0910.4905_arXiv.txt": { "abstract": "We use time-varying models of the coupled evolution of the HI, $\\rm H_2$ gas phases and stars in galaxy-sized numerical simulations to: a) test for the emergence of the Kennicutt-Schmidt (K-S) and the $\\rm H_2$-pressure relation, b) explore a realistic $\\rm H_2$-regulated star formation recipe which brings forth a neglected and potentially significant SF-regulating factor, and c) go beyond typical galactic environments (for which these galactic empirical relations are deduced) to explore the early evolution of very gas-rich galaxies. In this work we model low mass galaxies ($M_{\\rm baryon} \\le 10^9 \\msun$), while incorporating an independent treatment of CO formation and destruction, the most important tracer molecule of H2 in galaxies, along with that for the H2 gas itself. We find that both the K-S and the $\\rm H_2$-pressure empirical relations can robustly emerge in galaxies after a dynamic equilibrium sets in between the various ISM states, the stellar component and its feedback ($\\rm T\\ga 1\\,Gyr$). The only significant dependence of these relations seems to be for the CO-derived (and thus directly observable) ones, which show a strong dependance on the ISM metallicity. The $\\rm H_2$-regulated star formation recipe successfully reproduces the morphological and quantitative aspects of previous numerical models while doing away with the star formation efficiency parameter. Most of the $\\rm HI\\rightarrow H_2$ mass exchange is found taking place under highly non-equilibrium conditions necessitating a time-dependent treatment even in typical ISM environments. Our dynamic models indicate that the CO molecule can be a poor, non-linear, $\\rm H_2$ gas tracer. Finally, for early evolutionary stages ($\\rm T\\la 0.4\\,Gyr$) we find significant and systematic deviations of the true star formation from that expected from the K-S relation, which are especially pronounced and prolonged for metal-poor systems. The largest such deviations occur for the very gas-rich galaxies, where deviations of a factor $\\sim 3-4$ in global star formation rate can take place with respect to those expected from the CO-derived K-S relation. This is particularly important since gas rich systems at high redshifts could appear as having unusually high star-formation rates with respect to their CO-bright $\\rm H_2$ gas reservoirs. This points to a possibly serious deficiency of K-S relations as elements of the sub-grid physics of star formation in simulations of structure formation in the Early Universe. ", "introduction": "} In spite of the fact that the general character of the cycle through which galaxies convert their ISM to stars has been known for a long time, it has proven to be remarkably difficult to formulate a predictive theoretical framework for this process. Indeed, we know that throughout most of the Universe star formation takes place in molecular gas complexes \\citep[e.g.][]{Solomon2005, Omont2007}. The $\\rm HI\\rightarrow H_2$ phase transition is conditioned by a combination of sufficiently high HI column densities and pressures, and consequently star formation tends to concentrate in high density regions, e.g. in the central parts of galaxies, in spiral arms or high-pressure concentrations of gas formed by bulk gas motions or swept up by the shocks from supernovae and stellar winds of OB associations. The link between molecular gas and star formation is so tight that it has even been used to infer the distribution of the former by that of the latter, when the CO-$\\rm H_2$ conversion factor was still considered very uncertain \\citep{Rana1986}. In the Galaxy, where this link is best studied \\citep[e.g.][and references therein]{Blitz1997}, the latest results confirm star formation always taking place in CO-bright molecular clouds, even at very large galactocentric distances \\citep{Kobayashi2008}. In other galaxies this tight association has been verified in all cases where sufficient angular resolution is available \\citep[e.g.][]{Wong2002}. Thus it is fair to say that {\\it $\\rm H_2$ formation is a necessary prerequisite for star formation in galaxies,} and incorporating it in galaxy-sized numerical simulations of gas and stars is the single most important step currently missing from a realistic rendering of star formation in such models. Following the early and widespread observational evidence establishing the $\\rm H_2$-(star formation) link, the inclusion of the $\\rm H_2$ gas phase in numerical models of galaxies has occured only recently \\citep{Pelupessy2006, Dobbs2006, Robertson2008}, and the refinement of these models is an area of ongoing research. This is mainly due to the difficulty of tracking the dynamic and thermodynamic evolution of $\\rm H_2$ and its precursor phase, the Cold Neutral Medium HI \\citep[$\\rm n \\sim 5-100$ cm$^{-3}$, $\\rm T_k \\sim 60-200$K,][]{Wolfire2003} in galaxy-sized numerical models and due to the strong $\\rm H_2$ self-shielding complicating radiative transfer models of its far-UV radiation-induced destruction. The first problem has prevented most efforts from properly tracking the $\\rm HI\\rightarrow H_2$ phase transition in galaxies without resorting to simplifying steady-state solutions \\citep{Hidaka2002, Robertson2008} (suitable only for quiescent galactic environments), or to semi-empirical multiphase models \\citep[e.g.][]{Semelin2002} with limited predictive value. The second problem confounds even numerical simulations tracking the $\\rm HI\\rightarrow H_2$ phase transition in individual gas clouds where local approximations of the self-shielding HI/$\\rm H_2$ volume (necessary for numerically manageable solutions) can make the $\\rm H_2$ gas mass fraction a strong function of the chosen numerical resolution \\citep[e.g.][]{Glover2007a}. Finally a secondary, yet important problem of such single gas cloud simulations is posed by the constant boundary conditions assumed during their evolution, which are an unlikely setting for real gas clouds immersed in the ISM environment of a galaxy. In such environments cloud boundary conditions that powerfully influence the $\\rm HI\\rightarrow H_2$ phase transition, such as the ambient FUV radiation field and pressure, change on timescales comparable or shorter than ``internal'' cloud dynamic and thermodynamic timescales, especially in vigorously star forming environments \\citep[e.g.][]{Parravano2003, Wolfire2003, Pelupessy2006}. Despite the aforementioned difficulties the incorporation of the $\\rm H_2\\leftrightarrow HI$ gas phase interplay, and its strong role as star formation regulator, in numerical models holds the promise of large improvements in their handling of galaxy evolution, and the possibility of unveiling new, hitherto neglected, aspects of star formation feedback on the ISM. In this paper we apply our numerical models for the coupled evolution of gas (HI, $\\rm H_2$) and stars \\citep{Pelupessy2006} to the evolution of low mass galaxies ($M_{\\rm baryon}<10^9 \\msun$) in order to explore two new key directions, the first of which is the emergence of two important empirical relations found for galaxies in the local Universe: the Kennicutt-Schmidt (K-S) and the $\\rm H_2$--pressure relation. Secondly we will make a investigation of very gas-rich systems (more typical of the Early Universe), and check whether the aforementioned relations remain valid during their evolution. The latter is of crucial importance given the prominant role such empirical relations are given in describing the sub-grid star-formation/gas interplay in cosmological simulations of galaxy evolution (where the resolution limitations imposed by the simulation of large volumes preclude a detailed description of star formation). Finally along with our original time-dependent treatment of the $\\rm HI\\rightarrow H_2$ phase transition we also include the CO molecule, allowing direct comparisons to the {\\it observed} $\\rm H_2$ gas distributions, and a new independent investigation of the CO-$\\rm H_2$ relation within the dynamical setting of an evolving galaxy. The structure of our work is as follows: in section~\\ref{sec:model} we present relevant features of the model, show semi-analytical predictions for the $\\rm H_2$-pressure relation, and formulate the extention of the model so that includes CO, in section~\\ref{sec:sims} we present our detailed numerical simulations, and investigate the K-S, $\\rm H_2$--pressure and CO--$\\rm H_2$ empirical relations. In section~\\ref{sec:disc} we investigate and discuss the validity of the important K-S relation during early galaxy evolution stages, and for very gas-rich systems. We then summarize our conclusions in section~\\ref{sec:concl}. ", "conclusions": "\\label{sec:concl} We use our time-varying, galaxy-sized, numerical models of gas+stars that track the ISM thermodynamics and the $\\rm HI\\leftrightarrow H_2$ gas phase exchange, to investigate: a) the emergence of two prominent empirical relations deduced for galaxies in the local Universe: the Kennicutt-Schmidt (K-S) relation and the $\\rm H_2$-pressure relation, b) the effects of a more realistic $\\rm H_2$-regulated star formation recipe, and c) the evolution of very gas-rich systems. Our models now include a separate treatment for formation and destruction of the $\\rm H_2$-tracing CO molecule, which allows a direct comparison of such models with observations, and a new independent investigation of the CO-$\\rm H_2$ concomitance in the ISM of evolving galaxies. Our findings can be summarized as follows \\begin{itemize} \\item For ISM states of $\\rm H_2$/HI equilibrium, an $\\rm H_2$-pressure relation close to the one observed robustly emerges for a wide range of parameters, with a strong dependance mostly on metallicity. For the more realistic non-equilibrium $\\rm H_2$/HI states {\\it only the CO-bright $\\rm H_2$ phase shows an $\\rm H_2$-pressure relation similar to the one observed.} \\item The $\\rm H_2$-regulated star formation model successfully models star formation without the adhoc parameter of the local star formation efficiency adopted by most galaxy-sized numerical models, while incorporating a fundamental aspect of the star formation process. \\item A comparison between numerical models using the usual simple-delay (SD) and the new molecular-regulated (MR) star formation recipes reveals very few differences. It shows a factor of $\\sim 3-4$ more efficient star formation per CNM gas mass than the case of MR star formation. \\item We find little sensitivity of the global SF efficiency M(HI+$\\rm H_2$)/SFR to the SF recipe chosen, once dynamic equilibrium between ISM phases and stars is established, yielding confidence to (K-S)-type of relations emerging as a general characteristic of galaxies. \\item A non-equilibrium $\\rm HI\\leftrightarrow H_2$ gas mass exchange is revealed taking place under typical ISM conditions, demonstrating the need for a full dynamic rather than stationary treatment of these ISM phases. \\item The CO molecule can be a poor, non-linear, tracer of the true underlying $\\rm H_2$ gas distribution, especially in metal-poor systems, and even in those with very high gas mass fractions (more typically found at high redshifts). \\item A K-S relation robustly emerges from our time-dependent models, irrespective of the SF recipe used, after a dynamical equilibrium is established ($\\rm T\\ga $1\\,Gyr). The CO-derived K-S relation has a more shallow slope than the one involving the total gas mass, and as in the $\\rm H_2$-pressure relation, a strong dependance on metallicity is found. \\item At early evolutionary timescales ($\\rm T\\la 0.4\\,Gyr$) our models show {\\it significant and systematic deviations of the true star formation from that expected from the K-S relation,} which seem especially pronounced and prolonged for metal-poor systems. These deviations occur even for the CO-derived K-S relation (the more realistic one since CO is directly observable and traces the densest $\\rm H_2$ gas which ``fuels'' star formation), and even for metal-rich systems where CO tracks the $\\rm H_2$ gas well. \\item The largest deviations from the K-S relation occur at the earliest evolutionary stages of the systems modeled here ($\\rm T\\la 0.2 Gyr$) and for the most gas-rich ones. During this time significantly higher star formation rates per CO-bright $\\rm H_2$ gas mass occur, and such star-forming galaxies may have been already observed at high redshifts. \\end{itemize} Finally we must note that when it comes to the gas-rich galaxies accessible to current observational capabilities at high redshifts, our results, drawn for much less massive systems, remain provisional. Nevertheless for more massive gas-rich systems the larger amplitudes of ISM equilibrium-perturbing agents (e.g. SNs, far-UV radiation fields), and the shorter timescales that will characterize their variations are more likely than not to exaggerate the deviations of true star formation versus the one derived from (K-S)-type phenomenological relations. A dedicated observational effort to study such galaxies at high redshifts (soon to be dramatically enhanced by ALMA), as well as extending detailed numerical modeling of gas and stars to larger systems (as computational capabilities improve), can help establish whether (K-S)-type relations remain valid during most of the stellar mass built-up in galaxies, or only emerge after dynamic equilibrium has been reached during much latter evolutionary stages." }, "0910/0910.2242_arXiv.txt": { "abstract": "We present a spectroscopic sample of 910 distant halo stars from the Hypervelocity Star survey from which we derive the velocity dispersion profile of the Milky Way halo. The sample is a mix of 74\\% evolved horizontal branch stars and 26\\% blue stragglers. We estimate distances to the stars using observed colors, metallicities, and stellar evolution tracks. Our sample contains twice as many objects with $R>50$ kpc as previous surveys. We compute the velocity dispersion profile in two ways: with a parametric method based on a Milky Way potential model, and with a non-parametric method based on the caustic technique originally developed to measure galaxy cluster mass profiles. The resulting velocity dispersion profiles are remarkably consistent with those found by two independent surveys based on other stellar populations: the Milky Way halo exhibits a mean decline in radial velocity dispersion of $-0.38\\pm0.12$ \\kms\\ kpc$^{-1}$ over $15$50 kpc. Recently, \\citet{xue08} analyzed the Sloan Digital Sky Survey (SDSS) sample of 2,466 BHB stars and found a small decline in velocity dispersion with distance. Only 80 of the SDSS BHB stars are located at $R>50$ kpc. Here we use the distant halo stars from the hypervelocity star (HVS) program \\citep{brown05, brown06, brown06b, brown07a, brown07b, brown09a, brown09b} to measure the velocity dispersion profile of the Milky Way. Our dataset is a complete spectroscopic sample of 910 stars observed over 7300 deg$^2$ of the SDSS Data Release 6 imaging region. We gain increased leverage on the velocity dispersion profile for $R \\gtrsim 50$ kpc. We find an average decline in radial velocity dispersion of $0.38\\pm0.12$ \\kms\\ kpc$^{-1}$ over $1550$ kpc. The velocity dispersion profiles observed by these independent datasets are consistent at the 1.5-$\\sigma$ level, and have an average velocity dispersion slope identical to our result. Remarkably, no matter what tracers are used, observers find the same halo velocity dispersion profile. The velocity dispersion profile is a basis for measuring the total mass and mass distribution of the Milky Way halo. A companion paper by Gnedin et al.\\ (in preparation) presents the theoretical calculations that turn the observed velocity dispersion profile into a mass determination of the Milky Way. For further progress in measuring the Milky Way velocity dispersion profile, it is essential to identify tracers at distances $R>50$ kpc. It is difficult to find $R>50$ kpc tracers because of the steep decline in the density of the stellar halo. It is also very difficult for proper motion surveys to measure tracers at $R>50$ kpc distances. On-going spectroscopic radial velocity surveys, such as the SDSS-3 survey and our own HVS survey, promise to better trace the Milky Way in coming years." }, "0910/0910.5654_arXiv.txt": { "abstract": "We have invented a novel technique to measure the radio image of a pulsar scattered by the interstellar plasma with 0.1~mas resolution. We extend the ``secondary spectrum'' analysis of parabolic arcs by Stinebring et al.\\ (2001) to very long baseline interferometry and, when the scattering is anisotropic, we are able to map the scattered brightness astrometrically with much higher resolution than the diffractive limit of the interferometer. We employ this technique to measure an extremely anisotropic scattered image of the pulsar B0834$+$06 at 327~MHz. We find that the scattering occurs in a compact region about 420~pc from the Earth. This image has two components, both essentially linear and nearly parallel. The primary feature, which is about 16~AU long and less than 0.5~AU in width, is highly inhomogeneous on spatial scales as small as 0.05~AU. The second feature is much fainter and is displaced from the axis of the primary feature by about 9~AU. We find that the velocity of the scattering plasma is $16 \\pm 10$~\\kms approximately parallel to the axis of the linear feature. The origin of the observed anisotropy is unclear and we discuss two very different models. It could be, as has been assumed in earlier work, that the turbulence on spatial scales of ($\\sim 1000$~km) is homogeneous but anisotropic. However it may be that the turbulence on these scales is homogeneous and isotropic but the anisotropy is produced by highly elongated (filamentary) inhomogeneities of scale 0.05-16 AU. ", "introduction": "Radio pulsars provide a powerful tool for studying the ionized interstellar medium. The dispersion in their pulse arrival times probes the mean electron density. Their very small diameters ensure that they display the full range of scintillation and scattering phenomena, which probe the fine spatial structure in the electron density. Their pulse amplitudes exhibit a combination of diffractive and refractive intensity scintillation on times from seconds to months. Compact emission from some bright active galactic nuclei can also show scintillation on times of hours to months, albeit smoothed by the effect of their larger angular diameters. The large body of pulsar scintillation data has been interpreted in terms of homogeneous isotropic Kolmogorov turbulence in the interstellar plasma \\citep{ars95}. See reviews by \\citet{Ric90,Nar92}. However, a litany of observational evidence now points to the existence of compact ionized structures in the interstellar medium (ISM) whose scattering characteristics are well beyond those of such homogeneous isotropic Kolmogorov turbulence. In particular the scattering is seldom uniformly distributed along the line of sight. It is often dominated by one local region somewhere in the line of sight, which we refer to as a ``thin screen.'' Although it is unlikely to resemble a screen, it is thin with respect to the total line of sight from the source to the observer \\citep[e.g.,][]{putney,rjtr}. Inhomogeneity in the turbulence is required on kiloparsec scales to explain how the level of pulsar scattering varies with distance and Galactic coordinates \\citep{Cor91}. Inhomogeneity is also required on the parsec scale in the local ISM to explain the intermittent nature of the scintillation observed in a few quasars on hour-long time scales \\citep[e.g.,][]{DTdB03,KC06}. In addition, evidence for AU-scale inhomogeneity in the turbulence comes from extreme scattering events (ESEs), which are observed in a few quasars as rare, large (10 to 50\\%) variations in flux density over several weeks \\citep{Fiedler87,La01,Sen08}. These are generally seen as a decrease in flux density attributed to the passage across the line of sight of an ionized cloud which either scatters \\citep{Fiedler87} or refracts \\citep{Rom87} the radiation. In recent years there has also been increasing evidence that interstellar scattering (ISS) is not only inhomogeneous but also anisotropic. The most direct measure of anisotropy is through very long baseline interferometry (VLBI) imaging of scattered brightness, but it is only detectable on a few heavily scattered lines of sight \\citep[e.g.,][]{Des01}. Anisotropy in the ISS diffraction pattern has also been measured indirectly \\citep[e.g.,][]{Ric02, DTdB03,Col05,Big06}. Such observations give evidence for elongated fine structure in the ISM on scales of thousands of kilometers, suggesting anisotropic magneto-hydrodynamic turbulence controlled by the magnetic field as discussed by \\citet{Gol95} and \\citet{Spa99}. The examples cited above show departures from both homogeneity and isotropy in the ionized ISM. It is possible, but by no means proven, that the various phenomena have a common origin in a population of AU-scale anisotropic regions of enhanced density and turbulence which we here generically refer to as ``clouds.'' However, we note that such localized clouds must contain fine scale substructure that causes scattering at radio frequencies. Such a population of clouds presents a serious puzzle. Their number density must be many orders of magnitude greater than that of stars and their implied electron densities $n_{\\rm e} \\age 10$~cm$^{-3}$ are much higher than expected in pressure equilibrium in the warm ionized phase ISM \\citep{Rom87}. The discovery of parabolic arcs in the ISS of pulsars by \\citet{Sti01} has added a powerful new tool for probing such clouds. Many pulsars exhibit parabolic arcs in their secondary spectra (SS), which is the power spectrum (versus delay and Doppler frequency) of the dynamic spectrum of intensity (versus frequency and time). The arcs are sometimes narrow (in delay) which implies scattering by a thin layer. The distribution of power in the SS often reveals anisotropic scattering and in some cases there are discrete downwards facing ``arclets,'' which also imply scattering from isolated anisotropic clouds (``cloudlets''). See \\citet{Cor06,Wal04} for interpretation of the arcs. The SS allows a two dimensional reconstruction of the scattered image from observations at a single receiver, since each point in the SS isolates the scintillation power associated with interference between pairs of points on the scattering disk. However the reconstruction can be model dependent and has an inherent two-fold ambiguity \\citep{Cor06,Tra07}. Occasionally one can also see isolated peaks in the SS corresponding to narrow bandwidth fringes in the dynamic spectrum \\citep{Wol87,Ric97}. Such fringes result from interference between the normal primary (on-axis) scattering disk and the off-axis discrete cloud. Recent results from \\citet{Hil05} have compounded the difficulties in understanding the nature of the underlying ionized clouds. They observed the SS towards pulsar B0834$+$06 and found four distinct arclets scattered through 7 to 12~mas which they interpreted as originating from $\\sim$0.2~AU clouds requiring $n_{\\rm e} > 100$~cm$^{-3}$, similar to those invoked to explain ESEs towards quasars. By monitoring the evolution of structures in the secondary spectrum, they followed these clouds over three weeks and showed that they co-moved with the rest of the scattering material. In this paper we report VLBI observations of the scintillation from the same pulsar (B0834$+$06) in order to further investigate these clouds. We have developed a novel astrometry technique that makes use of SS-like quantities derived from the interferometer visibilities. Using these ``secondary cross spectra'' (defined in Table~\\ref{tab:defs}), we can accurately localize points on the scattering screen corresponding to high signal-to-noise pixels of the SS. We use the results of the astrometry of many such points to measure the distance and velocity of the interstellar clouds and so define a precise model for the scattering. Applying this precise model to the astrometric image with the scattering model allows us to eliminate all the otherwise troublesome ambiguities and validates the model. We then use this precise model to recover the scattered image with even greater angular resolution from the secondary spectrum itself. In our observations the Rayleigh resolution (i.e., synthesized beam) of the VLBI array is about 35~mas, the astrometric precision is about 1~mas, and structure in the scattered image recovered by modeling is found on a scale of $100$~$\\mu$as. ", "conclusions": "This paper describes a novel VLBI technique resulting in a two-dimensional image of the scattering screen of pulsar B0834$+$06. The baseband data that were recorded allowed high resolution dynamic spectra to be produced. The secondary spectra produced with the two-hour dynamic spectra could allow sharply defined arclets to be identified with delays as high as 1~ms. The scattered image was developed by astrometrically mapping points chosen from the secondary spectrum to bright points in the sky plane. These points were clustered in two clearly defined groups: a primary scattering disk which is elongated and inclined $27 \\pm 2^{\\circ}$ to the pulsar proper motion direction and a second, non-colinear, feature corresponding to the 1~ms feature of the secondary spectrum. Diagnostic measurements place the two features at essentially the same distance, 65\\% of the way to the pulsar. The two-dimensional distribution of points in the scattered image allows both transverse components of the effective velocity to be determined via relationships connecting the Doppler frequency with location in the image. The discrete feature at 1~ms delay contains about 4\\% of the total received power. This feature is expected to to be visible only for a few weeks during which time its delay should drift as the pulsar moves; the impact on timing this pulsar at $\\sim 327$~MHz due to such a feature is a time variable wander with magnitude $\\sim 40$~$\\mu$s. This should come as a caution to those aiming to perform precision pulsar timing at low frequencies on pulsars that exhibit the extreme forms of scintillation that are characteristic of B0834$+$06. Further, pulsars with sub-microsecond structure may experience apparent pulse profile evolution yielding additional complications in their timing. We were able to estimate the effective scintillation velocity vector, which depends on a distance-weigthed sum of the velocities of the pulsar, the Earth and the sacttering plasma. By using the published proper motion we estimated the velocity of the scattering plasma to be $16\\pm 10$~\\kms approximately parallel to the scattering axis. Since the errors in this interesting result are dominated by the uncertainty in the pulsar proper motion, we have undertaken a new set of VLBI measurements to improve its precision. The interpretation of ISS in pulsars has often assumed isotropy in the scattering. The extremely anisotropic scattering found here would substantially alter any quantitative modelling of the plasma were it to be a common feature in other regions on the interstellar medium. A description of the underlying plasma physics must await a resolution of the two possible geometries mentioned in the previous section, but the results and the method provide an exciting new glimpse of the ionized ISM at scales of 0.1 to 10~AU." }, "0910/0910.0071_arXiv.txt": { "abstract": "{In mean-field magnetohydrodynamics the mean electromotive force due to velocity and magnetic field fluctuations plays a crucial role. In general it consists of two parts, one independent of and another one proportional to the mean magnetic field. The first part may be nonzero only in the presence of mhd turbulence, maintained, e.g., by small-scale dynamo action. It corresponds to a battery, which lets a mean magnetic field grow from zero to a finite value. The second part, which covers, e.g., the $\\alpha$ effect, is important for large-scale dynamos. Only a few examples of the aforementioned first part of mean electromotive force have been discussed so far. It is shown that a mean electromotive force proportional to the mean fluid velocity, but independent of the mean magnetic field, may occur in an originally homogeneous isotropic mhd turbulence if there are nonzero correlations of velocity and electric current fluctuations or, what is equivalent, of vorticity and magnetic field fluctuations. This goes beyond the Yoshizawa effect, which consists in the occurrence of mean electromotive forces proportional to the mean vorticity or to the angular velocity defining the Coriolis force in a rotating frame and depends on the cross-helicity defined by the velocity and magnetic field fluctuations. Contributions to the mean electromotive force due to inhomogeneity of the turbulence are also considered. Possible consequences of the above and related findings for the generation of magnetic fields in cosmic bodies are discussed.} ", "introduction": "Mean--field magnetohydrodynamics has proved to be a useful tool for studying the behavior of mean magnetic fields in turbulently moving electrically conducting fluids (see, e.g., Moffatt 1979, Krause \\& R\\\"adler 1980, Brandenburg \\& Subramanian 2005). Within this framework both the magnetic field $\\bB$ and the fluid velocity $\\bU$ are split into mean parts, $\\bmB$ and $\\bmU$, and fluctuating parts, $\\bb$ and $\\bu$. Starting from the induction equation governing $\\bB$ it is concluded that the mean magnetic field $\\bmB$ has to obey \\EQ \\p_t \\bmB = \\eta \\bnab^2 \\bmB + \\bnab \\x (\\bmU \\x \\bmB + \\bscE ) \\, , \\quad \\bnab \\cdot \\bmB = 0 \\, . \\label{eq01} \\EN Here, $\\eta$ means the magnetic diffusivity of the fluid, for simplicity considered as independent of position, and $\\bscE$ the mean electromotive force caused by the velocity and magnetic fluctuations, \\EQ \\bscE = \\lan \\bu \\x \\bb \\rangle \\, . \\label{eq03} \\EN Mean fields are defined by some kind of averaging satisfying the Reynolds rules. They are denoted either by overbars or synonymously by angle brackets. The induction equation governing $\\bB$ also implies \\EQA && \\!\\!\\!\\!\\!\\!\\!\\!\\!\\!\\! \\p_t \\bb = \\eta \\bnab^2 \\bb + \\bnab \\x [ (\\bu \\x \\bb)' + \\bmU \\x \\bb + \\bu \\x \\bmB], \\nonumber\\\\ && \\qquad \\qquad \\qquad \\qquad \\qquad \\bnab \\cdot \\bb = 0 \\, , \\label{eq05} \\ENA where $(\\bu \\x \\bb)' = \\bu \\x \\bb - \\lan \\bu \\x \\bb \\ran$. With this in mind we may conclude that $\\bscE$ can be represented as a sum of two parts, \\EQ \\bscE = \\bscE^{(0)} + \\bscE^{(\\mB)} \\, , \\label{eq07} \\EN where $\\bscE^{(0)}$ is a functional of $\\bu$ and $\\bmU$, and $\\bscE^{(\\mB)}$ a functional of $\\bu$, $\\bmU$ and $\\bmB$, which is linear in $\\bmB$ but vanishes if $\\bmB$ is zero everywhere and at all past times (see, e.g., R\\\"adler 1976, 2000, R\\\"adler \\& Rheinhardt 2007). These statements apply independently on whether or not $\\bu$ or $\\bmU$ depend on $\\bmB$. If they depend on $\\bmB$ and the total variation of $\\bscE$ with $\\bmB$ is considered, $\\bscE^{(0)}$ may well depend on $\\bmB$, and $\\bscE^{(\\mB)}$ need not be linear in $\\bmB$. A non--zero $\\bscE^{(0)}$ corresponds to a battery. Assume for a moment that equation (\\ref{eq01}) for $\\bmB$ with $\\bscE = \\bzo$ has no growing solutions. If then $\\bscE^{(0)}$ takes non--zero values, but $\\bscE^{(\\mB)}$ remains equal to zero, $\\bmB$ grows, even if initially equal to zero, to a finite magnitude determined by $\\bscE^{(0)}$. If, on the other hand, $\\bscE^{(0)}$ remains equal to zero a non--zero $\\bscE^{(\\mB)}$ may allow (if it has a suitable structure) a dynamo, that is, let an arbitrarily small seed magnetic field $\\bmB$ grow exponentially (in the absence of back--reaction on the fluid motion even endlessly). A small non--zero $\\bscE^{(0)}$ may deliver a seed field for such a dynamo. This possibility has been already discussed in the context of young galaxies (Brandenburg \\& Urpin 1998). In most of the general representations and applications of mean--field magnetohydrodynamics the part $\\bscE^{(0)}$ of the electromotive force $\\bscE$ has been ignored. Indeed, if it occurs at all, it decays to zero in the course of time except in cases in which an independent magnetohydrodynamic turbulence exists, e.g., as a result of a small--scale dynamo. The possibility of a non--zero $\\bscE^{(0)}$ due to local, that is small--scale, dynamos in the solar convection zone has been discussed by R\\\"adler (1976). We express his statements here by \\EQ \\bscE^{(0)} = c_\\gamma \\bgamma + c_\\Omega \\bOmega + c_{\\gamma \\Omega} \\bgamma \\x \\bOmega \\, , \\label{eq09} \\EN where $\\bgamma$ is a gradient, e.g., of the turbulence intensity, $\\bOmega$ the angular velocity responsible for the Coriolis force, and $c_\\gamma$, $c_\\Omega$ and $c_{\\gamma \\Omega}$ some coefficients. More precisely, $c_\\gamma$ and $c_{\\gamma \\Omega}$ are scalars and $c_\\Omega$ is a pseudoscalar. Another interesting result has been derived by Yoshizawa (1990, see also Yoshizawa 1993 or Yoshizawa, Itoh \\& Itoh 2003). Considering an originally homogeneous isotropic magnetohydrodynamic turbulence under the influence of a mean flow or a rigid--body rotation, or both, he found \\EQ \\bscE^{(0)} = c_W \\bW + c_\\Omega \\bOmega \\, , \\label{eq11} \\EN where $\\bW = \\bnab \\x \\bmU$ is the mean vorticity, $\\bOmega$ again the angular velocity responsible for the Coriolis force, and $c_W$ and $c_\\Omega$ are pseudoscalars coefficients which are, roughly speaking, proportional to the cross--helicity $\\lan \\bu \\cdot \\bb \\ran$. This result has recently been used for an interpretation of the Archontis dynamo (Sur \\& Brandenburg 2009). The main purpose of this paper is to demonstrate that a mean electromotive force $\\bscE^{(0)}$ proportional to the mean fluid velocity $\\bmU$ may occur in originally homogeneous isotropic magnetohydrodynamic turbulence. This should be expected as soon as there is a non--zero correlation between the fluctuating parts of velocity and electric current, $\\bu$ and $\\bj = \\mu_0^{-1} \\bnab \\x \\bb$, or, what is equivalent, between the fluctuating parts of vorticity and magnetic field, $\\bomega = \\bnab \\x \\bu$ and $\\bb$; as usual, $\\mu_0$ means the magnetic permeability. We express this condition roughly by saying that $\\lan \\bu \\cdot \\bj \\ran$ or $\\lan \\bomega \\cdot \\bb \\ran$ have to be unequal to zero. Unlike $\\lan \\bu \\cdot \\bb \\ran$, which characterizes the linkage between vortex tubes and magnetic flux tubes, $\\lan \\bu \\cdot \\bj \\ran$ quantifies the linkage between vortex tubes and current tubes. In Section \\ref{sec2} we explain the basis of our calculations and provide general relations for the determination of the mean electromotive force $\\bscE^{(0)}$. In Section \\ref{sec3} we derive results for homogeneous isotropic turbulence, in particular the last--mentioned one, and we also reproduce that given by (\\ref{eq11}). Proceeding then in Section \\ref{sec4} to inhomogeneous turbulence we report on results related to those indicated in (\\ref{eq09}). The relevance of the results obtained in this paper and the need of further work are discussed in Section \\ref{sec5}. ", "conclusions": "\\label{sec5} The most remarkable result of our calculations is that the mean electromotive force $\\bscE^{(0)}$ in a homogeneous isotropic magnetohydrodynamic turbulence may have a contribution proportional to the mean fluid velocity $\\bmU$, that is, $\\bscE^{(0)} = c_U \\, \\bmU + \\cdots$. We have labeled the occurrence of this contribution as $\\lan \\bu \\cdot \\bj \\ran$ effect. The coefficient $c_U$ turned out to be in general unequal to zero if only a non--zero correlation exists between the fluctuating parts of the fluid velocity and the electric current density, $\\bu$ and $\\bj = \\mu_0^{-1} \\bnab \\x \\bb$, or between the fluctuating parts of the vorticity and the magnetic field, $\\bomega = \\bnab \\x \\bu$ and $\\bb$. As far as the second--order correlation approximation applies, $c_U$ vanishes for $\\eta = \\nu$, that is, it changes its sign if $\\nu / \\eta$ varies and passes through $\\nu / \\eta = 1$. The occurrence of magnetohydrodynamic turbulence does not automatically imply non--zero correlations of $\\bu$ and $\\bj$, or $\\bomega$ and $\\bb$. It depends on the special circumstances whether, e.g., $\\lan \\bu \\cdot \\bj \\ran$ or $\\lan \\bomega \\cdot \\bb \\ran$ are different from zero and what their signs are. For this and other reasons more work is needed to explore the importance of the $\\lan \\bu \\cdot \\bj \\ran$ effect in specific settings. In general the $\\lan \\bu \\cdot \\bj \\ran$ effect is accompanied by the $\\lan \\bu \\cdot \\bb \\ran$ effects. It has then also to be investigated which of these effects dominates. At first glance there seems to be a inconsistency of our result concerning the $\\lan \\bu \\cdot \\bj \\ran$ effect in so far as $\\bscE^{(0)}$ should not depend on the choice of the frame of reference but $\\bmU$ obviously does. We must however keep in mind that we have fixed the frame of reference in our calculation by assuming that there isotropic turbulence occurs in the limit $\\bmU \\to \\bzo$. When estimating the $\\lan \\bu \\cdot \\bj \\ran$ effect we have therefore to specify $\\bmU$ as the mean velocity of the fluid relative to the frame in which the assumed causes of turbulence (in simulations the forcing) would, in the absence of mean motion, just lead to isotropic turbulence. There is still another issue which has to be considered when applying our result concerning the $\\lan \\bu \\cdot \\bj \\ran$ effect to a specific situation. The deviation of the turbulence from isotropy due to the homogeneous part of the mean motion is crucial for that effect. It has to be scrutinized whether such a deviation indeed occurs under the considered circumstances. The sometimes assumed ``Galilean invariance\" of the turbulence (e.g., Sridhar \\& Subramanian 2009), that is, its independence of that part of the mean motion, would exclude the $\\lan \\bu \\cdot \\bj \\ran$ effect. In principle we could have determined $\\bscE^{(0)}$ in a frame in which $\\bmU$ vanishes. An anisotropy of the turbulence, as calculated in the frame used above, would lead to a non--vanishing contribution $\\bscE^{(00)}$ instead of $c_U \\bmU$, which then has to be considered as another description of the $\\lan \\bu \\cdot \\bj \\ran$ effect." }, "0910/0910.0840_arXiv.txt": { "abstract": "We report the results of a study of the rest-frame ultraviolet (UV) spectrum of the Cosmic Eye (J213512.73$-$010143), a luminous ($L \\sim 2 L^\\ast$) Lyman break galaxy at $z_{\\rm sys}=3.07331$ magnified by a factor of $\\sim 25$ via gravitational lensing by foreground mass concentrations at $z = 0.73$ and 0.33. The spectrum, recorded at high resolution and signal-to-noise ratio with the ESI spectrograph on the Keck\\,{\\sc ii} telescope, is rich in absorption features from the gas and massive stars in this galaxy. The interstellar absorption lines are resolved into two components of approximately equal strength and each spanning several hundred \\kms\\ in velocity. One component has a net blueshift of $-70$ \\kms\\ relative to the stars and H\\,{\\sc ii} regions and presumably arises in a galaxy-scale outflow similar to those seen in most star-forming galaxies at $z = 2$--3. The other is more unusual in showing a mean \\textit{redshift} of $+350$ \\kms\\ relative to $z_{\\rm sys}$; possible interpretations include a merging clump, or material ejected by a previous star formation episode and now falling back onto the galaxy, or more simply a chance alignment with a foreground galaxy. In the metal absorption lines, both components only partially cover the OB stars against which they are being viewed. However, there must also be more pervasive diffuse gas to account for the near-total covering fraction of the strong damped \\lya\\ line, indicative of a column density $N$(H\\,{\\sc i})\\,$ = (3.0 \\pm 0.8) \\times 10^{21}$\\,cm$^{-2}$. We tentatively associate this neutral gas with the redshifted component, and propose that it provides the dust `foreground screen' responsible for the low ratio of far-infrared to UV luminosities of the Cosmic Eye. The C\\,{\\sc iv} P~Cygni line in the stellar spectrum is consistent with continuous star formation with a Salpeter initial mass function, stellar masses from 5 to 100\\,\\msun, and a metallicity $Z \\sim 0.4 Z_{\\odot}$. Compared to other well-studied examples of strongly lensed galaxies, we find that the young stellar population of the Cosmic Eye is essentially indistinguishable from those of the Cosmic Horseshoe and MS\\,1512-cB58. On the other hand, the interstellar spectra of all three galaxies are markedly different, attesting to the real complexity of the interplay between starbursts and ambient interstellar matter in young galaxies observed during the epoch when cosmic star formation was at its peak. ", "introduction": "\\label{sec:introduction} High redshift star-forming galaxies are often identified by a break in their ultraviolet continuum that is due to the Lyman limit, partly from interstellar (within the galaxy) and primarily from intergalactic H\\,{\\sc i} absorption below 912\\,\\AA\\ (Steidel et al. 1996). Since the discovery of these `Lyman break galaxies' (or LBGs), samples of galaxies at $z = 2$--4 have increased a thousand-fold. Despite the variety of methods now employed to select high redshift galaxies, LBGs remain the most common and most extensively studied class of such objects. Detailed studies of the physical properties of \\textit{individual} galaxies have in general been limited by their faintness ($m_{\\cal R}^{\\ast} = 24.4$, Steidel et al. 1999; Reddy et al. 2008). In a few cases, however, gravitational lensing by foreground massive galaxies, groups, or clusters has afforded rare insights into the stellar populations and interstellar media of galaxies at $z = 2$--4 (e.g. Pettini et al. 2000, 2002; Teplitz et al. 2000; Lemoine-Busserolle et al. 2003; Smail et al. 2007; Swinbank et al. 2007; Cabanac, Valls-Gabaud, \\& Lidman 2008; Siana et al. 2008, 2009; Finkelstein et al. 2009; Hainline et al. 2009; Quider et al. 2009; Swinbank et al. 2009; Yuan \\& Kewley 2009 and references therein). The order-of-magnitude boost in flux provided by strong lensing makes it possible to record the spectra of these galaxies with a combination of high resolution and high signal-to-noise ratio (S/N) which for normal, unlensed galaxies will have to wait until the next generation of 30+\\,m optical/infrared telescopes. Focusing on the rest-frame UV spectra in particular, our group has so far published observations of two of these strongly lensed high redshift galaxies: MS\\,1512-cB58 and J1148+1930 (cB58 and the `Cosmic Horseshoe' respectively; Pettini et al. 2002; Quider et al. 2009). Both are $\\sim L^{*}$ galaxies at redshifts $z\\sim 2.5$ magnified by a factor of $\\sim 30$ (Seitz et al. 1998; Dye et al. 2008). A considerable amount of data that are inaccessible to typical low-resolution studies is contained in these rest-UV spectra, including information on interstellar gas composition and kinematics, on the initial mass function (IMF) of starbursts at high $z$, and clues to the geometry of the gas and stars. One of the motivations for this work is to assess the range of properties possessed by high redshift star-forming galaxies. By comparing the UV spectra of cB58 and the Cosmic Horseshoe, Quider et al. (2009) showed that the young stellar populations of these galaxies are very similar, as are the metallicities and kinematics of their interstellar media. On the other hand, the two galaxies exhibit clear differences in the covering fractions of their stars by the interstellar gas, differences which are reflected in the strikingly different morphologies of their Lyman~$\\alpha$ lines (a damped absorption profile in cB58 and a double-peaked emission profile in the Horseshoe). Studies at other wavelengths (Teplitz et al. 2000; Hainline et al. 2009) have also shown that cB58 and the Cosmic Horseshoe have comparable star formation rates, $\\rm{SFR} \\sim 50$--100\\,\\msun~yr$^{-1}$, at the upper end of the range of values measured in star-forming galaxies at $z = 2$--3, and similar dynamical masses, $M_{\\rm dyn} \\sim 1 \\times 10^{10}$\\,\\msun, typical of luminous galaxies at this epoch (Pettini et al. 2001; Erb et al. 2006b,c; Reddy et al. 2006). Clearly, several examples of strongly lensed galaxies need to be studied at such levels of detail for a full characterization of galaxy properties at these redshifts. Fortunately, many new cases have been discovered recently, mostly from dedicated searches in imaging data from the Sloan Digital Sky Survey (Estrada et al. 2007; Belokurov et al. 2007, 2009; Ofek et al. 2008; Shin et al. 2008; Kubo et al. 2009; Lin et al. 2009; Wen et al. 2009). Other surveys have capitalised on the superior spatial resolution of \\textit{Hubble Space Telescope (HST)} to identify high redshift galaxies strongly lensed into multiple images or arcs; the $z = 3.0733$ galaxy which is the focus of the present paper, dubbed the `Cosmic Eye', was indeed discovered by Smail et al. (2007) from an \\textit{HST} Snapshot program targeting high luminosity X-ray clusters. In this paper we present high resolution observations of the rest-frame UV spectrum of the Cosmic Eye. The paper is organized as follows. In Section~\\ref{sec:eyesummary} we summarize the known properties of the Cosmic Eye from previous studies at a variety of wavelengths, while Section~\\ref{sec:observations} has details of our observations and data reduction. We analyze the interstellar spectrum of the Cosmic Eye in Section~\\ref{sec:ISM}, and the composite spectrum of its young stellar population in Section~\\ref{sec:stars}. Section~\\ref{sec:intervening} focuses on the numerous intervening absorption systems that potentially confound our interpretation of the galaxy's intrinsic features. In Section~\\ref{sec:discussion} we discuss the implications of our findings; finally, we summarize our main conclusions in Section~\\ref{sec:conclusions}. Throughout the paper, we adopt a cosmology with $\\Omega_{\\rm M} = 0.3$, $\\Omega_{\\Lambda} = 0.7$, and $H_0 = 70$\\,km~s$^{-1}$~Mpc$^{-1}$. ", "conclusions": "\\label{sec:conclusions} Strong gravitational lensing of a $z = 3.07331$ star-forming [${\\rm SFR} \\simeq 50\\,M_\\odot$~yr$^{-1}$ for a Chabrier (2003) IMF] galaxy magnifies it by a factor of $\\sim 25$ and distorts its image into two $3^{\\prime\\prime}$ long arcs which have been collectively named the `Cosmic Eye' by their discoverers, Smail et al. (2007). The high magnification has allowed us to use the ESI spectrograph on the Keck\\,{\\sc ii} telescope to record the galaxy's rest-frame UV spectrum with high resolution and S/N ratio. By analyzing these data together with existing observations of the Cosmic Eye at other wavelengths, we have reached the following main conclusions. (i) The interstellar absorption lines exhibit two components, of approximately equal strength, which are respectively blueshifted by $-70$\\,\\kms\\ and redshifted by $+350$\\,\\kms\\ relative to the stars and H\\,{\\sc ii} regions. While these values apply to the gas with the highest apparent optical depths, both components include absorption spanning several hundred \\kms. We associate the blueshifted component with a galaxy-wide outflow similar to, but possibly weaker than, those seen in most star-forming galaxies at $z = 2$--3. The redshifted absorption is very unusual, and may represent gas ejected by a previous episode of star formation and now falling back onto the galaxy, or a merger viewed along a favourable line of sight. Alternatively, it may just be a chance superposition of another galaxy along the line of sight. (ii) Both components of the metal absorption lines show indications that they do not fully cover the OB stars against which they are being viewed; we estimate covering fractions of $\\sim 70$\\% and $\\sim 85$\\% for the blueshifted and redshifted component respectively. There must also be more pervasive diffuse gas, because the strong damped \\lya\\ line, corresponding to a column density $N$(H\\,{\\sc i})\\,$= (3.0 \\pm 0.8) \\times 10^{21}$\\,cm$^{-2}$, covers at least 95\\% of the UV stellar continuum. We tentatively associate this high column density of H\\,{\\sc i} with the redshifted component of the metal lines, where absorption from ionized species is weak or missing altogether, and propose that it provides the `foreground screen' of dust responsible for the lower-than-expected far-infrared luminosity of the Cosmic Eye found by Siana et al. (2009) with the \\textit{Spitzer Space Telescope}. (iii) The internal kinematics of the galaxy lensed into the Cosmic Eye are very complex, our data now adding outflow, and possibly even inflow, to the rotation and velocity dispersion already known from the integral field spectroscopy with adaptive optics by Stark et al. (2008). Ordered rotation, chaotic motions, and outflow all seem to be of comparable magnitude, with $v_{\\rm rot} \\sin i \\approx \\sigma_0 \\approx v_{\\rm blue} \\simeq 50$--70\\,\\kms. However, we do not have a model yet of how these different motions fit together into one coherent kinematic picture. (iv) Turning to the stellar spectrum, we find that the C\\,{\\sc iv} P~Cygni profile is well fit by a \\textit{Starburst99} stellar population model spectrum having continous star formation with a Salpeter IMF, stellar masses from 5 to 100\\,\\msun, and a LMC/SMC metallicity of $Z \\sim 0.4$\\,\\Zsun. The P~Cygni profiles of the Cosmic Eye, the Cosmic Horseshoe, and MS\\,1512-cB58, three high redshift star-forming galaxies studied at high spectral resolution, are all nearly identical. This is not unexpected, however, when we consider that in each case we see the integrated light of several hundred thousand O-type stars, and that these three galaxies have similar metallicities and dynamical timescales over which star formation is taking place. (v) The metallicity $Z \\simeq 0.4 Z_\\odot$ deduced for the O stars in the Cosmic Eye is lower by a factor of $\\sim 2$ than the value derived from the analysis of the strong nebular lines from its H\\,{\\sc ii} regions using the R23 index. We consider this apparent discrepancy to reflect systematic offsets between different abundance estimators, rather than intrinsic inhomogeneities in the chemical composition of stars and ionized gas. (vi) The interpretation of both interstellar and stellar features in the UV spectrum of the Cosmic Eye is complicated by the presence of numerous intervening absorption lines associated with eight absorption systems at redshifts $z_{\\rm abs} = 2.4563$--3.0528. These narrow features are not resolved in existing low-resolution data, highlighting the caution that should be exercised in interpreting the spectra that are typically available for unlensed LBGs. In closing, the new data presented here further emphasize the complexity of the physical conditions which prevailed in actively star-forming galaxies at redshifts $z = 2$--3. It is remarkable that the three strongly-lensed galaxies targeted by our ESI observations, while showing very similar young stellar populations, are all different in the detailed properties of their interstellar media. Whether such variety is simply the result of different geometries and viewing angles, or has its roots in more fundamental physical reasons remains to be established. Fortunately, with the increasing attention being given to detailed studies of gravitationally lensed galaxies, we can look forward with optimism to a more comprehensive empirical picture of galaxy formation coming together in the years ahead." }, "0910/0910.5462_arXiv.txt": { "abstract": "We comment on the calculation mistake in the paper ``$w$ and $w'$ of scalar field models of dark energy'' by Takeshi Chiba, where $w$ is the dark energy equation of state and $w'$ is the time derivative of $w$ in units of the Hubble time. The author made a mistake while rewriting the phantom equation of motion, which led to an incorrect generic bound for the phantom model and an incorrect bound for the tracker phantom model on the $w\\textrm{--}w'$ plane. ", "introduction": "\\subsection{Generic bound} The phantom model is scalar field dark energy having negative kinetic term and its equation of state $w<-1$. The energy density and the pressure are given by $\\rho=-\\dot\\phi^2/2+V$ and $p=-\\dot\\phi^2/2-V$, respectively. The equation of motion is given by \\begin{equation}\\label{q1} \\ddot\\phi+3H\\dot\\phi-V_{,\\phi}=0 \\ . \\end{equation} Therefore, the phantom field tends to roll up the potential $V$. The phantom equation of motion can be rewritten as~\\cite{Kujat:2006vj} \\begin{equation}\\label{q2} \\pm {V_{,\\phi}\\over V}=\\sqrt{{-3\\kappa^2(1+w)\\over \\Omega_{\\phi}}} \\left(1+{x'\\over 6}\\right) \\ , \\end{equation} where the plus sign corresponds to $\\dot\\phi>0$ and the minus sign to the opposite, $\\kappa^2=8\\pi G$ and $\\Omega_{\\phi}$ is the fractional energy density of the phantom field. The variable $x$ is defined as \\begin{equation}\\label{q3} x=\\ln\\left(-{1+w\\over 1-w}\\right) \\ , \\end{equation} and $x'$ is the derivative of $x$ with respect to $\\ln a$, which is related with $w'$ as \\begin{equation}\\label{q4} x'={2w'\\over (1-w)(1+w)} \\ . \\end{equation} Since the left-hand side of Eq.~(\\ref{q2}) is positive for the up-rolling phantom field, we have $1+x'/6>0$. Therefore, using Eq.~(\\ref{q4}), the upper bound on $w'$ is obtained as \\begin{equation}\\label{q5} w'<-3(1-w)(1+w). \\end{equation} Note that the upper bound on $w'$ for phantom is smoothly connected to the lower bound on $w'$ for quintessence, the latter of which is shown as Eq.~(2.6) in~\\cite{Chiba:2005tj}. In~\\cite{Chiba:2005tj}, there is a calculation mistake that the rewritten equation of motion is obtained as \\begin{equation}\\label{q6} \\mp {V_{,\\phi}\\over V}=\\sqrt{{-3\\kappa^2(1+w)\\over \\Omega_{\\phi}}} \\left(-1+{x'\\over 6}\\right) \\ . \\end{equation} As a consequence, an incorrect bound on $w'$ is obtained as \\begin{equation}\\label{q7} w'>3(1-w)(1+w). \\end{equation} \\subsection{Tracker phantom}\\label{TP} Following the approach in~\\cite{Chiba:2005tj}, the bound can be tightened for the tracker field. Taking the derivative of Eq.~(\\ref{q2}) with respect to $\\phi$, we obtain the tracker equation for the phantom field \\begin{eqnarray}\\label{q8} \\Gamma-1 &=& \\frac{3(w_B-w)(1-\\Omega_{\\phi})}{(1+w)(6+x')}-\\frac{(1-w)x'} { 2(1+w)(6+x')} \\nonumber \\\\ & \\ & -\\frac{2x''}{(1+w)(6+x')^2} \\ , \\end{eqnarray} where $\\Gamma=VV_{,\\phi\\phi}/V_{,\\phi}^2$, $w_B$ is the equation of state of the background matter, and $x''$ is the second derivative of $x$ with respect to $\\ln a$. Therefore, for the tracker solution where $w$ is nearly constant and $\\Omega_{\\phi}$ is initially negligible, $w$ is given by \\begin{equation}\\label{q9} w={w_B-2(\\Gamma -1)\\over 2(\\Gamma -1) +1} \\ . \\end{equation} Thus $\\Gamma<1/2$ is required for tracker phantom, which has $w<-1$. Following~\\cite{Chiba:2005tj}, we consider a solution in which initially $w$ follows the tracker solution in Eq.~(\\ref{q9}) and then evolves toward $-1$. Therefore, the tracker $w$ in Eq.~(\\ref{q9}) is a lower bound of $w$. In such solution, $x'$ eventually stops decreasing and then increases back to a value near zero. The minimum value of $x'$, $x_m'$, gives an upper bound on $w'$ via Eq.(\\ref{q4}). To find $x_m'$, we put $x''=0$ and $w_B=0$ in Eq.~(\\ref{q8}) and find that \\begin{equation}\\label{q10} x_m'=-6{w(1-\\Omega_{\\phi})+2(1+w)(\\Gamma -1)\\over (1-w)+2(1+w)(\\Gamma -1)} \\ . \\end{equation} Since $x_m'$ is an increasing function of $w$, a lower bound on $x_m'$ is given by that of $w$, for which we take the tracker $w$ in Eq.~(\\ref{q9}) and obtain \\begin{equation}\\label{q11} x_m'>\\frac{6w\\Omega_{\\phi}}{1-2w}>\\frac{6w}{1-2w} \\ . \\end{equation} With Eq.~(\\ref{q4}), we then obtain the upper bound on $w'$ \\begin{equation}\\label{q12} w'<\\frac{3w(1-w)(1+w)}{1-2w} \\ . \\end{equation} In \\cite{Chiba:2005tj}, there is a mistake that the tracker equation for the phantom field is given as \\begin{eqnarray}\\label{q13} \\Gamma-1 &=& \\frac{3(w_B-w)(1-\\Omega_{\\phi})}{(1+w)(6-x')}-\\frac{(1-w)x'} { 2(1+w)(6-x')} \\nonumber \\\\ & \\ & +\\frac{2x''}{(1+w)(6-x')^2} \\ . \\end{eqnarray} As a consequence, the resulting $x_m'$ is \\begin{eqnarray}\\label{q14} x_m'&=&-6{w(1-\\Omega_{\\phi})+2(1+w)(\\Gamma -1)\\over (1-w)-2(1+w)(\\Gamma -1)} \\\\&>&6w\\Omega_{\\phi}>6w. \\end{eqnarray} Therefore, an incorrect bound on $w'$ is given as \\begin{equation}\\label{q15} w'<3w(1-w)(1+w). \\end{equation} The bounds for the phantom model from \\cite{Chiba:2005tj} are shown in Fig.~\\ref{fig1} and the corrected ones we obtain are shown in Fig.~\\ref{fig2}. \\begin{figure} \\includegraphics[width=8.3cm]{origional} \\caption{\\label{fig1}Bounds on $w'$ as a function of $w$ for the phantom model from \\cite{Chiba:2005tj}. The solid curve is the generic lower bound of the phantom. The dashed curve is the upper bound of the tracker phantom. The allowed region for the tracker phantom is filled with the grey color. } \\end{figure} \\begin{figure} \\includegraphics[width=8.3cm]{corrected} \\caption{\\label{fig2}Corrected bounds we obtain on $w'$ as a function of $w$ for the phantom model. The solid curve is the generic upper bound of the phantom. The dashed curve is the upper bound of the tracker phantom. The allowed region for the tracker phantom is filled with the grey color.} \\end{figure} ", "conclusions": "" }, "0910/0910.2418_arXiv.txt": { "abstract": "We apply a reflection-based model to the best available {\\it XMM-Newton} spectra of X-ray bright UltraLuminous X-ray (ULX) sources (NGC~1313 X--1, NGC~1313 X--2, M~81 X--6, Holmberg~IX X--1, NGC~5408 X--1 and Holmberg~II X--1). A spectral drop is apparent in the data of all the sources at energies 6--7\\,keV. The drop is interpreted here in terms of relativistically-blurred ionized reflection from the accretion disk. A soft-excess is also detected from these sources (as usually found in the spectra of AGN), with emission from O K and Fe L, in the case of NGC~5408 X--1 and Holmberg~II X--1, which can be understood as features arising from reflection of the disk. Remarkably, ionized disk reflection and the associated powerlaw continuum provide a good description of the broad-band spectrum, including the soft-excess. There is no requirement for thermal emission from the inner disk in the description of the spectra. The black holes of these systems must then be highly spinning, with a spin close to the maximum rate of a maximal spinning black hole. The results require the action of strong light bending in these sources. We suggest that they could be strongly accreting black holes in which most of the energy is extracted from the flow magnetically and released above the disc thereby avoiding the conventional Eddington limit. ", "introduction": "Ultra-luminous X-ray (ULX) sources are point like, nonnuclear sources observed at other galaxies, with observed luminosities greater than the Eddington luminosity for a $10\\,{\\rm M}_{\\odot}$ stellar mass black hole (BH), with $L_{X}{\\ge}10^{39}$\\,${\\rm erg\\,s^{-1}}$ \\citep{f1}. The true nature of these objects is still open to debate \\citep{mc1}. One fundamental issue is whether the emission is isotropic or beamed along our line-of-sight. A possible scenario for geometrical beaming involves super-Eddington accretion during phases of thermal-timescale mass transfer \\citep{k1}. Alternatively, if the emission is isotropic and the Eddington limit is not violated, ULX must be fuelled by accretion onto Intermediate-Mass BH (IMBH), with masses in the range 100-10\\,000 ${\\rm M}_{\\odot}$ \\citep{cm1}. Currently, there is no agreement regarding the nature of these sources. It is possible that some ULX appear very luminous because of a combination of moderately high mass, mild beaming and mild super-Eddington emission. It is also possible that ULX are an inhomogeneous population, comprised of both a subsample of IMBH and moderately beamed stellar mass black holes \\citep{f2,mc1}. NGC~5408 X--1, HOLM~II X--1, M 81 X--9 (which is in the companion-dwarf galaxy Holmberg IX, hereafter called HOLM~IX X--1), the ULX in NGC~1313 and M 81~X--6 are ULX located in dwarf (NGC~5408 X--1, HOLM~II X--1 and HOLM~IX X--1) and spiral (NGC~1313 X--1, NGC~1313 X--2 and M 81 X--6) galaxies, respectively. All these ULX peak in X-ray luminosity above $L_{X}=1{\\times}10^{40}$\\,${\\rm erg\\,s^{-1}}$, thus being excellent targets for testing spectral models to the data. They are nearby and located at distances of $D=4.8,3.5,3.7,3.63,4.50$\\,Mpc (\\citealt{karachentsev02,stobbart06,liu08}), for NGC~5408 X--1, HOLM~IX X--1, NGC~1313 X--1, M~81 X--6 and HOLM~II X--1 respectively. NGC~5408 X--1 is among the few ULX for which (10-200\\,mHz) Quasi-Periodic Oscillations (QPO) have been found (\\citealt{stroh1,dewangan06}). NGC~5408~X-1 exhibits X-ray timing and spectral properties analogous to those exhibited by Galactic stellar-mass black hole in the {\\it very high} or {\\it steep power-law} state \\citep{remi1}, but with the characteristic variability timescales (QPO and break frequencies) consistently scaled down \\citep{stroh1}. For NGC~5408 X--1 the inferred characteristic size for the X-ray emitting region is ${\\approx}$ 100 times larger than the typical inner disk radii of Galactic black holes. The highest signal-to-noise spectra of ULX can often be fitted by composite models of a thermal disk together with a hard-powerlaw tail, with a low measured disk temperature of ${\\approx}0.2$\\,keV (\\citealt{miller03,miller04}). If such factors are entirely due to a higher black hole mass, they imply an IMBH with $M{\\approx}100-10\\,000$\\,${\\rm M}_{\\odot}$. Both the morphology and flux of the optical high-excitation nebulae detected around some ULX probably rule out strong beaming as the origin of the X-ray emission (\\citealt{pakull03,kaaret04,soria1,abolmasov07}). The detection for most ULX of spectral curvature, in the form of a deficit of photons at energies $E{\\ge}2$\\,keV (\\citealt{roberts05,stobbart06,miyawaki09}) has led to the suggestion that most ULX have spectral properties that do not correspond to any of the accretion states known in Black Hole Binaries (BHB), making it unlikely that ULX are powered by sub-Eddington flows onto an IMBH \\citep{roberts07}. The application of Comptonization models to the data (\\citealt{stobbart06,gladstone09} and references therein) results in strikingly high and low values for the coronal opacity (${\\tau}{\\ge}5$) and the electron temperature ($kT_{e}=1-3$\\,keV), difficult to explain for the expected physical conditions in a corona surrounding the black hole. This is very different to the typical values found for BHB during the {\\it low/hard} state, with spectra dominated by Comptonization \\footnote{It has to be noted that similar values, i.e. $kT_{e}=1-3$\\,keV and ${\\tau}{\\ge}5$, have been found in the description of the spectra of BHB during the {\\it steep powerlaw} state (e.g. GRO~J1655--40 and GRS~1915+105; \\citealt{makishima00,kubota04,ueda09}).}, and appears irreconcilable with the IMBH model, which assumes that they operate as simple scaled-up BHB. Here, we present an alternative interpretation of the spectral shape, based on a physically-justified model commonly used on other accreting black holes. The soft part of the spectrum ($E{\\le}1$\\,keV) -- the soft X-ray excess -- and the high-energy curvature are just aspects of a reflection spectrum expected from accretion (\\citealt{guilbert88,george91}). A major component of reflection, the broad iron K line, has been found in many Seyfert galaxies (\\citealt{tanaka95,nandra08}), accreting stellar-mass black holes (\\citealt{miller07,reis09a}) and even accreting neutron stars (\\citealt{cackett08,reis09b}). Both the soft excess and the relativistic broad iron K line have recently been demonstrated to be part of the same physical process, i.e., the reaction of the disk to irradiation from a high-energy source, in the Seyfert-1 galaxy 1H 0707-495 \\citep{fabian09}. In this paper we investigate whether reflection models can account for the spectra of these ULX and, if so, we put them in the context of the accreting black holes known so far. For this study, we choose the ULX with the best available data and with the longest exposure time observations (${\\approx}100$\\,ks of exposure time) of the {\\it XMM-Newton} satellite. By using only the highest quality data from the widest band pass, highest sensitivity instruments available we expect to make good statements on the accretion processes in these ULX. In Section \\ref{observ} we describe the observations and data used, in Section \\ref{spec_anal} we report on the results of the application of fits with the reflection model to the spectra and in Section \\ref{discuss} discuss the results obtained. ", "conclusions": "\\label{discuss} Remarkably, a relativistically-blurred, reflection-dominated model gives a very good fit to the sample of ULX with best quality (100\\,ks of time exposure) {\\it XMM-Newton} data. Very skewed iron line profiles have been found, implying that the emission region is very close to the black-hole and this fact allowed us to determine their spin. We find that all the ULX in our sample are close to maximally rotating, with the exception of NGC~5408 X--1 and HOLM~II X--1, for which the steep powerlaw did not allow us to properly determine parameters from the Fe line. We have found that, the sources showing a {\\it reflection dominated} spectrum (NGC~1313 X--1, NGC~1313 X--2 and M~81 X--6) have maximally spinning black-holes. This fact can be understood if light bending is an important effect for these sources. For these sources, the high-energy source (i.e. the jet or corona) is very close to the black hole and the observed direct continuum (i.e. powerlaw) is very low. This would correspond to {\\tt Regime I} of \\citet{miniutti04}, corresponding to a low height of the primary source, where strong light suffered by the primary radiation dramatically reduces the observed powerlaw emission component at infinity. HOLM~IX X--1 shows a {\\it powerlaw dominated} spectrum with a slightly narrower Fe line profile. This would correspond to {\\tt Regime II} of \\citet{miniutti04}, when the height of the primary source is larger, and the observed direct continuum (i.e. powerlaw) stronger. The line profile is narrower than in regime I because the emissivity profile of the disk is flatter, as result of a nearly isotropic illumination. For NGC~5408 X--1 and HOLM~II X--1, we find a clear reflection component but the steepness of the spectra means that the Fe K features are less well-measured. The spectral solution we have found here for ULX suggests a new explanation for accretion onto spinning black holes. Our model assumes that the dominant source of radiation is a power-law continuum produced a few gravitational radii above a rapidly spinning black hole. Little thermal radiation is produced by the disc and is undetected by the current data. We assume that the power for this source is extracted magnetically from the disc and transferred to the emission region by magnetic fields. Some of the power may even be extracted magnetically from the spin of the black hole \\citep{blandford77}. Since radiation is only produced above the disc, radiation pressure need not oppose accretion. Indeed it will help squash the disc and maintain the high surface density required for our relatively low ionization parameters and thus observable reflection. (Note that strong light bending occurs in the region being considered which is very close to a black hole, meaning that there is beaming but in the {\\emph opposite} sense to that normally envisaged.) The relevant radiation for computing the physical Eddington limit in this situation is the thermal disc radiation, not the power-law continuum and reflection associated with it. Since we detect no such thermal radiation, then the situation may be sub-Eddington, even for stellar mass black holes. Whether this solution can work depends on the extent to which magnetic energy extraction can be clean, in the sense of not requiring considerable thermal energy release. We note that the accretion flow is super-Eddington in the conventional interpretation in which the total energy release is considered (especially since some radiation falls straight into the black hole). If the above solution is appropriate for these objects then they can either be stellar mass black holes with masses of ${\\approx}10\\Msun$ or IMBH of $100$s$\\Msun$. Their spin is then likely to be natal. If they form through stellar collapse and the progenitor star was highly spinning, then it would be likely that the black hole would remain highly spinning. Many studies have argued that a binary companion can spin up a massive star but the magnitude of the spin up is still a matter of debate (e.g., \\citealt{paczynski98,fryer99,brown00}). Accretion would presumably come from a binary companion. An interesting alternative is that they accrete from a debris disc produced by fallback from the (failed) supernova or collapsar \\citep{li03}. This could account for the high metallicity we infer (note that hydrogen-poor accretion would also slightly raise the Eddington limit). In summary, some ULX may consist of fast spinning black holes accreting rapidly from a metal-rich disc. Accretion power passes magnetically to the primary emission region which lies just outside the accretion flow. Strong reflection occurs, creating a soft excess and very broad iron-K features in the 0.3-10 keV X-ray spectrum. Sensitive hard X-ray spectra in the 10--50~keV band may test this scenario." }, "0910/0910.4714_arXiv.txt": { "abstract": "We study the prospects for detecting neutrino masses from the galaxy angular power spectrum in photometric redshift shells of the Dark Energy Survey (DES) over a volume of $\\sim 20\\, h^{-3}$ Gpc$^3$, combined with the Cosmic Microwave Background (CMB) angular fluctuations expected to be measured from the Planck satellite. We find that for a $\\Lambda$-CDM concordance model with 7 free parameters in addition to a fiducial neutrino mass of $M_{\\nu} = 0.24$ eV, we recover from DES\\&Planck the correct value with uncertainty of $ \\pm 0.12$ eV (95 \\% CL), assuming perfect knowledge of the galaxy biasing. If the fiducial total mass is close to zero, then the upper limit is $0.11$ eV(95 \\% CL). This upper limit from DES\\&Planck is over 3 times tighter than using Planck alone, as DES breaks the parameter degeneracies in a CMB-only analysis. The analysis utlilizes spherical harmonics up to 300, averaged in bin of 10 to mimic the DES sky coverage. The results are similar if we supplement DES bands (grizY) with the VISTA Hemisphere Survey (VHS) near infrared band (JHK). The result is robust to uncertainties in non-linear fluctuations and redshift distortions. However, the result is sensitive to the assumed galaxy biasing schemes and it requires accurate prior knowledge of the biasing. To summarize, if the total neutrino mass in nature greater than 0.1eV, we should be able to detect it with DES\\&Planck, a result with great importance to fundamental Physics. ", "introduction": "\\renewcommand{\\thefootnote}{\\fnsymbol{footnote}} \\setcounter{footnote}{1} \\footnotetext{E-mail: lahav@star.ucl.ac.uk} Neutrinos are so far the only dark matter candidates that we actually know exist. It is now established from solar, atmospheric, reactor and accelerator neutrino experiments that neutrinos have non-zero mass, but their absolute masses are still unknown. Cosmology could provide an upper limit on the sum of neutrino masses \\citep[for review see e.g.][]{elg05,lesg06}. The growth of Fourier modes with comoving wavenumber $k > k_{\\rm nr}$ will be suppressed because of neutrino free-streaming, where \\begin{equation} k_{\\rm nr} = 0.026\\left(\\frac{m_{\\nu}}{1\\;{\\rm eV}}\\right)^{1/2} \\Omega_{\\rm m}^{1/2}\\;h\\,{\\rm Mpc}^{-1}, \\label{eq:knr} \\end{equation} for three equal-mass neutrinos, each with mass $m_\\nu$. The current mass upper limit, obtained using Cosmic Microwave Background (CMB) WMAP5 data, SN Ia and the BAO from 2dFGRS and SDSS, is $M_{\\nu} \\equiv \\sum m_{\\nu} < 0.61$ eV at 95\\% CL \\citep{koma09}. The challenge now is to bring down reliably the upper limits to the 0.1 eV level or even detect the neutrino mass. In this way Cosmology could resolve the mass scale of neutrinos. The new generation of deep wide surveys can play a key role in setting a tight upper limit on the neutrino mass, and possibly detect it if the true neutrino mass is sufficiently high. This is due to the order of magnitude increase in volume of the new surveys. Here we study specifically the ability to set an upper limit on the neutrino mass from the galaxy clustering expected in the photometric redshift survey Dark Energy Survey (DES), combined with Planck CMB measurements. The reason a survey like DES would be effective is its large volume, $\\sim 20\\, h^{-3}$ Gpc$^3$, and large number of galaxies, $\\sim 300$ million. Crudely, for a spectroscopic survey, where accurate redshifts are known, the error on the power spectrum scales with the survey effective volume $V_{\\rm eff}$ as \\begin{equation} \\Delta P(k)/P(k) \\propto 1/{\\sqrt {V_{\\rm eff} }}. \\label{eq:dpkveff} \\end{equation} On the other hand, the suppression is proportional in the linear regime to $f_{\\nu} = \\Omega_{\\nu}/\\Omega_m$ \\citep{hu98,kia08}, where \\begin{equation} \\Omega_{\\nu} = \\frac{\\Sigma_i m_i}{93.14 h^2 ~ \\rm eV}. \\label{eq:onumass} \\end{equation} We expect therefore (when all other cosmological parameters are fixed) that the determination of the upper limit on the neutrino mass would be inversely proportional to ${\\sqrt {V_{\\rm eff} } }$. From the 2dF Galaxy spectroscopic redshift survey, covering a volume of roughly 0.2 $(h^{-1} Gpc)^3$, the upper limit on the sum of neutrino mass is about 2 eV at 95 \\% CL \\citep{elg02}. Had DES been a spectroscopic survey with volume of about 20 $(h^{-1} Gpc)^3$, i.e. 100 times larger, we would expect an upper limit of 0.2 eV on the sum of neutrino mass. Our detailed calculation below yields an upper limit of 0.1eV for DES\\&Planck. This is tighter than the above back-of-the-envelope calculation, probably as Planck priors are incorporated, and the effective volumes above are only given crudely. However, DES is photometric redshift survey, so the radial component of distance to galaxies is significantly degraded, resulting in a poorer estimate of the power spectrum \\citep[e.g.][]{blk05}. Therefore we prefer in this analysis to quantify the galaxy clustering as angular (spherical harmonic) $C_{\\rm \\ell}$ power spectrum derived in photometric redshift shells which are wide enough relative to the photometric redshift errors, and to derive the resulting upper limits more carefully and quantitatively. We defer the comparison of $P(k)$ and $C_{\\rm \\ell}$ approaches to future studies. The utility of photometric redshifts is now well-established, with many successful techniques being employed \\citep[e.g.][]{coll04,abd09}. The cosmological parameter constraints resulting from future photometric redshift imaging surveys have been simulated by several authors \\citep[e.g.][]{seo03,dol04,zhan06}. Application to data such as the SDSS LRG samples were given by \\cite{blk07} and \\cite{pad07}. These studies mainly emphasized the detection of baryon acoustic oscillations in the galaxy clustering pattern. Apart from the specific application to DES, the present paper illustrates more generally the determination of neutrino mass from photo-z surveys, to our knowledge for the first time. This paper is organized as follows. In Sections \\ref{sec:galsur} we summarize the DES and VHS surveys and Planck, in Section \\ref{sec:photz} we present the photometric redshifts for DES and DES\\& the Vista Hemisphere Survey (VHS) combined filters. Sections \\ref{sec:clform} give the formalism for the galaxy angular power spectrum, and the associated joint likelihood with Planck. Section \\ref{sec:result} presents the results for the basic observational and theoretical scenarios, while Section \\ref{sec:analys} provides extensions of the analysis. An overall discussion is given in Section \\ref{sec:conclu}. ", "conclusions": "\\label{sec:conclu} We study the prospects for detecting neutrino masses from the galaxy angular power spectrum in photo-z shells in the Dark Energy Survey, combined with CMB fluctuations as will be measured by Planck. Although the core science case for DES is Dark Energy, we see that DES can provide us with other important extra science, such as neutrino mass. Our main conclusions are: \\begin{itemize} \\item We forecast for DES\\&Planck a 2-sigma error of total neutrino mass $\\Delta M_{\\nu} \\approx 0.12$ eV. If the true neutrino mass is very close to zero, then we can obtain an upper limit of 0.11 eV (95\\% CL). \\item This upper limit from DES+Planck is over 3 times tighter than using Planck alone, as DES breaks the parameter degeneracies in a CMB-only analysis. \\item The results are sensitive to the assumed galaxy biasing, and stand if the galaxy bias in known to within 2 per cent. This is feasible given other analyses of the galaxy bias such as the three point correlation function \\citep{ross07}. \\item The results are robust to uncertainties in non-linear fluctuations and redshift distortion. \\item The results are similar if we supplement DES bands (grizY) with the VISTA Hemisphere Survey (VHS) near infrared band (JHK). \\end{itemize} DES can also be used to extract information on neutrino mass via other techniques, e.g. weak gravitational lensing, as considered recently for other imaging surveys \\citep{kit08, ichi09}. We note that the level of sensitivity for neutrino mass from DES\\& Planck is of much relevance for comparison with the direct measurement of the neutrino mass from laboratory experiments. E.g. the KATRIN tritium beta decay experiment. Furthermore, the DES \\& Planck measurements can be combined with laboratory experiments to derive more accurate neutrino masses \\citep{host07}." }, "0910/0910.3302_arXiv.txt": { "abstract": "We present a hard X-ray spectrum of unprecedented quality of the Galactic supernova remnant W49B obtained with the {\\it Suzaku} satellite. The spectrum exhibits an unusual structure consisting of a saw-edged bump above 8~keV. This bump cannot be explained by any combination of high-temperature plasmas in ionization equilibrium. We firmly conclude that this bump is caused by the strong radiative recombination continuum (RRC) of iron, detected for the first time in a supernova remnant. The electron temperature derived from the bremsstrahlung continuum shape and the slope of the RRC is $\\sim$1.5~keV. On the other hand, the ionization temperature derived from the observed intensity ratios between the RRC and K$\\alpha$ lines of iron is $\\sim$2.7~keV. These results indicate that the plasma is in a highly overionized state. Volume emission measures independently determined from the fluxes of the thermal and RRC components are consistent with each other, suggesting the same origin of these components. ", "introduction": "\\label{sec:introduction} W49B (G43.3-0.2) is a Galactic supernova remnant (SNR) with strong X-ray line emissions from highly ionized atoms. It exhibits centrally filled X-rays inside a bright radio shell with a radius of 100 arcsec (Pye et al. 1984). The distance to W49B remains uncertain. It was estimated to lie at a distance of $\\sim$8~kpc (Radhakrishnan et al. 1972; Moffett \\& Reynolds 1994). Using the spectral distribution of HI absorption, however, Brogan \\& Troland (2001) found no clear evidence that W49B is closer to the sun than W49A, which is located at a distance of $\\sim$11.4~kpc (Gwinn et al. 1992). This may indicate a possible association of W49B with the star-forming region W49A. The near-infrared narrowband images indicate a barrel-shaped structure with coaxial rings, which is suggestive of bipolar wind structures surrounding massive stars (Keohane et al. 2007). They also showed an X-ray image from {\\it Chandra}, which has a jet-like structure along the axis of the barrel. They interpreted these findings as evidence that W49B had exploded inside a wind-blown bubble in a dense molecular cloud. Using {\\it ASCA} data, Hwang et al. (2000) showed that broadband modeling of the remnant's spectrum required two thermal components (0.2~keV and 2~keV) and significant overabundances of Si, S, Ar, Ca, Fe, and Ni. They confirmed that most of the X-ray emitting plasma was nearly in collisional ionization equilibrium (CIE). They also found evidence for Cr and Mn K$\\alpha$ emission. Kawasaki et al.\\ (2005) claimed the presence of ``overionized\" plasma in W49B through the analysis of {\\it ASCA} 2.75--6.0~keV spectrum. They measured intensity ratios of the H-like K$\\alpha$ (hereafter Ly$\\alpha$) to He-like K$\\alpha$ (He$\\alpha$) lines of Ar and Ca to obtain the ionization temperature ($kT_z$), and found that $kT_z$ ($\\sim$2.5~keV) is significantly higher than the electron temperature ($kT_{\\rm e}$) determined from the bremsstrahlung continuum shape ($\\sim$1.8~keV). Miceli et al.\\ (2006) adopted the same analysis procedure for the {\\it XMM-Newton} spectrum of the central region but found no evidence for the overionized state. In this letter, we report the firm evidence for overionized plasma in W49B using data from the X-ray Imaging Spectrometers (XIS: Koyama et al.\\ 2007) onboard the {\\it Suzaku} satellite (Mitsuda et al.\\ 2007). \\begin{figure}[t] \\includegraphics[scale=.60]{f1.eps} \\caption{Vignetting-corrected XIS image of W49B in the 1.5--7~keV band shown on a linear intensity scale. Data from the three active XISs are combined. Gray contours indicate every 10\\% intensity level relative to the peak surface brightness. The red circle and orange annulus indicate the source and background regions, respectively. The XIS field of view is shown by the black square. \\label{fig:image}} \\end{figure} ", "conclusions": "\\label{sec:discussion} We have found, for the first time, the strong He-RRC of Fe from W49B. Yamaguchi et al.\\ (2009) recently discovered the H-RRC of Si and S from a middle-aged SNR, IC~443, and hence this is the second discovery of a clear RRC from an SNR. We also discovered RRC-accompanied recombination lines, which may provide good diagnostics for the overionized plasma. We review a quantitative verification of our spectral analysis and discuss the implications of the results. \\subsection{Contribution of the Recombination Lines} We discuss the validity of the best-fit fluxes of the RRC and recombination lines in Table~1. The recombination cross section of the H-like ions into a level of $n$ is approximately given as shown below (e.g., Nakayama \\& Masai \\ 2001). \\begin{equation} \\sigma_{n} \\propto \\frac{1}{n^3} \\bigg(\\frac{3}{2} \\frac{kT_{\\rm e}}{E_{\\rm edge}}+ \\frac{1}{n^2} \\bigg) ^{-1} \\label{eq:RC-line} \\end{equation} We apply this approximation for the He-like ions, but $\\sigma_{1}$ must be reduced by half because one electron is already at the ground state. In principle, the recombination line flux can be estimated by the branching ratio to various levels, but these processes are very complicated. We, therefore, base our discussion only on a simple picture. We compare the predicted capture and observed transition rates normalized with the $n$ = 1 value. We can estimate the capture rates using Equation~2 as $\\sigma_2/\\sigma_1$ = 0.62, $\\sigma_3/\\sigma_1$ = 0.25, and $(\\sigma_4+\\sigma_5+...)/\\sigma_1$ = 0.34. On the other hand, the observed flux rates are He$\\alpha_{\\rm rec}$/He-RRC = 0.47 ($\\pm$0.10), He$\\beta_{\\rm rec}$/He-RRC = 0.053 ($\\pm$0.023), and He$\\gamma$-$\\infty_{\\rm rec}$/He-RRC = 0.19 ($\\pm$0.03).\\footnote{Throughout this paper, all the errors in the parentheses are at 90\\% confidence level.} By taking the fractions between the observed and predicted rates, we obtain 76 ($\\pm$16)\\%, 21 ($\\pm$9)\\%, and 56 ($\\pm$9)\\% for the He$\\alpha_{\\rm rec}$, He$\\beta_{\\rm rec}$, and He$\\gamma$-$\\infty_{\\rm rec}$ lines, respectively. These fractions for the He$\\beta_{\\rm rec}$ and He$\\gamma$-$\\infty_{\\rm rec}$ lines may be conceivable, because one electron at $n$ = 1 suppresses direct transitions from excited levels ($n \\ge$ 3$\\to$1) for the He-like ions. The fraction of He$\\alpha_{\\rm rec}$ is slightly smaller than expected because there should be a significant contribution of the cascade decay electrons from higher levels ($n \\ge$ 3$\\to$2$\\to$1). The real He$\\alpha_{\\rm rec}$ flux may be somewhat larger due to the uncertainty of the response function, but the contribution of the He$\\alpha_{\\rm rec}$ flux relative to the total He$\\alpha$ is only $\\lesssim$10\\% and does not affect results and following discussion. \\subsection{Electron and Ionization Temperatures} The electron temperatures determined by the bremsstrahlung continuum shape and the RRC slope are 1.52 (1.50--1.53) keV and 1.43 (1.30--1.59) keV, respectively. These consistent results indicate a common origin of these emissions. The ionization temperatures directly reflect the ion fractions and hence can be determined as shown below. From the best-fit model in Table~1, the flux ratios of Ly$\\alpha$/He$\\alpha$, He-RRC/He$\\alpha$, and H-RRC/He-RRC are given as 0.016 (0.014--0.018), 0.12 (0.11--0.13), and 0.023 ($\\le$0.10). In Figure~4, we compare these values with the modeled emissivity ratios derived from the radiation code of Masai (1994) for a plasma of $kT_{\\rm e}$ = 1.5~keV. We obtain $kT_z$ = 2.58 (2.46--2.68)~keV, 2.68 (2.63--2.73) keV, and 2.55 ($\\le$3.65) keV, respectively, for the above ratios. We also compare the Ly$\\alpha$/He$\\alpha$ ratio with that derived from the APEC code (Smith et al. 2001), although this code is valid only for a CIE state. We obtain $kT_z$ = 2.46 (2.39--2.54) keV, which is within the margin of error of the above results. The ionization temperatures ($\\sim$2.7~keV) are significantly higher than the electron temperatures ($\\sim$1.5~keV), indicating that the plasma is in a highly overionized state. The first hint of overionized plasma in W49B was found by Kawasaki et al. (2005). Although they analyzed different elements (Ar and Ca) in different energy bands (2.75--6.0~keV), and derived $kT_z$ from the Ly$\\alpha$/He$\\alpha$ ratio using the CIE plasma code, their results ($kT_z$ $\\sim$2.5~keV and $kT_{\\rm e}$ $\\sim$1.8~keV) are nearly consistent with ours. Our claim is more essential because it is based on clear detection of the recombination structures. \\subsection{Iron and Nickel Abundances} The abundances listed in Table~1 are valid only for the CIE state and should be modified in the overionized case. Since no plasma code can be applied to the overionized plasma currently, we make possible modifications using an available APEC code. The He$\\alpha$ intensity is proportional to $Z\\times \\epsilon(kT_{\\rm e}, kT_z)$, where $Z$ and $\\epsilon(kT_{\\rm e}, kT_z)$ are the abundance of the element (solar) and the total emissivity for the He$\\alpha$ for $kT_{\\rm e}$ and $kT_z$ (cm$^3$s$^{-1}$), respectively. The emissivities of the He-, Li-, Be-, and B-like ions for Fe and the He- and Li-like ions for Ni are modified by multiplying the ion-fraction ratio between $kT_z$ = 2.7~keV and 1.5~keV (Mazzotta et al. 1998). The total emissivity is given by adding those in individual ionization states. Multiplying by $\\epsilon$(1.5~keV, 1.5~keV)/$\\epsilon$(1.5~keV, 2.7~keV), we obtain the real abundances in the overionized state as $Z_{\\rm Fe}$$\\sim$4.9~solar and $Z_{\\rm Ni}$$\\sim$5.2~solar. Both the elements are highly over abundant, indicating an ejecta origin of the plasma. \\subsection{Volume Emission Measure} To check the consistency of the common origin of the bremsstrahlung and RRC emissions, we compare the volume emission measure (VEM). The VEM is given by $\\int n_e n_{\\rm H} dV/(4\\pi D^2)$, where $n_e$, $n_{\\rm H}$, $V$, and $D$ are the electron and hydrogen densities (cm$^{-3}$), emitting volume (cm$^3$), and distance to the source (cm), respectively. The VEM of the VAPEC component ($VEM_{\\rm VAPEC}$) is derived from Table~1 as $VEM_{\\rm VAPEC}$ = 1.61 (1.60 -- 1.62) $\\times 10^{13}$ cm$^{-5}$. On the other hand, the VEM of the RRC plasma ($VEM_{\\rm RRC}$) is calculated from \\begin{equation} F_{\\rm RRC}=\\alpha _1(kT_{\\rm e}) \\times \\frac{n_{\\rm Fe}}{n_{\\rm H}} \\times \\kappa_{\\rm H-like}(kT_z) \\times VEM_{\\rm RRC}, \\label{eq:rrcflux} \\end{equation} where $F_{\\rm RRC}$, $\\alpha _1(kT_{\\rm e})$, $n_{\\rm Fe}$, and $\\kappa_{\\rm H-like}(kT_z)$ are the flux of the He-RRC (cm$^{-2}$s$^{-1}$), the K-shell recombination rate coefficient for $kT_{\\rm e}$ (cm$^{3}$s$^{-1}$), the number density of Fe (cm$^{-3}$), and the ion fraction of H-like Fe for $kT_z$, respectively. According to Badnell (2006), the total radiative recombination rate of He-like Fe at $kT_{\\rm e}$ = 1.5~keV is given as $\\sim$3.9$\\times 10^{-12}$ cm$^{3}$s$^{-1}$. The recombination rate into the ground state is given using Equation~2 as $\\sigma_{1}/(\\sigma_1 + \\sigma_2 +...)\\sim$0.45. The value of $n_{\\rm Fe}/n_{\\rm H}$ is calculated using $Z_{\\rm Fe}$ in Section 4.3 and the number density of the solar photosphere (Anders \\& Grevesse 1989). Using the observed He-RRC flux, we obtain\\footnote{Here, we consider that the $kT_z$ error is 2.4--2.8~keV. The large error of $VEM_{\\rm RRC}$ is due to this effect.} $VEM_{\\rm RRC}$ = 0.81 (0.58 -- 1.60)$\\times 10^{13}$ cm$^{-5}$. The two independent estimations of VEM give consistent results, supporting the same origin of the overionized plasma in W49B. The origin of the overionized plasma in SNRs is an open question, and beyond the scope of this paper. We simply note the possibility of the cooling of electrons via thermal conduction (Kawasaki et al. 2005) or a more drastic cooling caused when the blast wave breaks out of some ambient matter into the rarefied interstellar medium (Yamaguchi et al. 2009). If the latter is the case, a massive progenitor that had blown a stellar wind is likely to be favored. \\begin{figure}[t] \\includegraphics[scale=.60]{f4.eps} \\caption{Predicted emissivity ratios of Ly$\\alpha$/He$\\alpha$ (black), He-RRC/He$\\alpha$ (red), and H-RRC/He-RRC (green) of Fe as a function of the ionization temperature ($kT_z$) for an electron temperature of 1.5~keV (Masai 1994). The horizontal dashed lines represent 90\\% errors of the observed values. \\label{fig:ratio}} \\end{figure}" }, "0910/0910.1594_arXiv.txt": { "abstract": "{% The formation of early-type dwarf (dE) galaxies, the most numerous objects in clusters, is believed to be closely connected to the physical processes that drive galaxy cluster evolution, like galaxy harassment and ram-pressure stripping. However, the actual significance of each mechanism for building the observed cluster dE population is yet unknown. Several distinct dE subclasses were identified, which show significant differences in their shape, stellar content, and distribution within the cluster. Does this diversity imply that dEs originate from various formation channels? Does ``cosmological'' formation play a role as well? I try to touch on these questions in this brief overview of dEs in galaxy clusters. } ", "introduction": "\\begin{figure} \\centering \\includegraphics[width=50mm]{dEbigsmall_2_s.eps} \\caption{Two Virgo cluster dEs (VCC\\,856 \\& 839) in a deep exposure with ESO 2.2m/WFI. The image shows an area of $6'\\times5'$. } \\label{dEbigsmallfig} \\end{figure} Early-type dwarf (dE) galaxies play a key role in understanding galaxy cluster evolution. Their low mass and low density make them more susceptible to physical effects than giant galaxies, and they are a large population, outnumbering all other galaxy types in dense environments by far. They are thus ideal probes of the mechanisms that govern galaxy formation and evolution in a cluster environment. Nevertheless, dEs are also interesting in their own right. Despite their rather unspectacular appearance, a surprising complexity in their characteristics has become evident, in terms of kinematics (rotating vs.\\ pressure supported), structure (flat vs.\\ round), and spatial distribution (cluster center vs.\\ outskirts). This zoo of dEs and their possible origin(s) is an ongoing challenge for observers and theorists. ", "conclusions": "\\subsection{Early versus late formation} Most dEs in cluster regions of high density are pressure supported (Fig.~\\ref{rotfig}), whereas in intermediate and low-density regions, more galaxies seem to be flattened by rotation or be rotationally supported\\linebreak \\citep{vanZee2004a,Toloba2009}. Does this, by itself, tell us something about the formation history of those dEs? Simulations show that relatively flat low-mass cluster galaxies experiencing repeated strong tidal interactions become dynamically hotter: their $v/\\sigma$ decreases and they become ``rounder'', i.e.\\ their axial ratio increases\\linebreak \\citep{Mastropietro2005}. Such interactions are more frequent in the central region of a cluster. Thus, if all dEs once had similar levels of rotational support, those in the central cluster region would have become dynamically hotter, in agreement with the observational findings (I call this {\\it case 1}). Likewise, if those that are now closer to the center have resided in the cluster for a longer time, they experienced more such interactions, leading to the same result ({\\it case 2}). Similar scenarios apply to the observation that galaxies in the central cluster regions have higher stellar population ages on average \\citep{Smith2008}: ram-pressure stripping, and also tidal stripping, are stronger in the central region, thus quenching star formation earlier in the galaxies that reside there. But without further considerations, it cannot be decided whether those galaxies have spent a longer time in the cluster, and were therefore stripped earlier (case 2), or whether they were simply stripped more efficiently (case 1). From case 2, one might be tempted to conclude a later infall and transformation of the progenitors of those dEs that are now in the cluster outskirts and are significantly flattened by rotation. However, case 1 does not imply anything about when the dEs, or their progenitors, have entered the cluster, or whether they have always been part of the cluster since its formation phase. This is the point at which input from the theoretical side is essential: for case 2, we need to have an understanding of the timescale between the infall of a group of progenitor galaxies and the point when these galaxies (or their remnants) reach a stage in which they are concentrated towards the cluster center and show a peaked and fairly regular distribution of heliocentric velocities. For case 1, probably even more complex input physics is needed, since semi-analytical model predictions for ``cosmological dwarfs'' that form in dense environments need to be tested against the observations. Equally important are observational studies of dEs at significantly earlier epochs, and comparisons to the local population (cf.\\ \\citealt{Andreon2008}; \\citealt{Harsono2009}, also this issue). \\subsection{Multiple origins?} Would it be possible, despite the complex characteristics and several different subclasses, to explain all dEs by just a single formation channel? One could, for example, imagine that dEs with disks (dE(di)s) do not form an intrinsically different subclass, but instead they simply constitute the flat tail of the other dEs, with their disk components having not been destroyed yet. The fraction of dE(di)s with nuclei ($\\sim$75\\%, \\citealt{p4}) agrees with the overall fraction of nucleated dEs in the (bright) magnitude range of the dE(di)s. However, while bright non-nucleated dEs (without disks or blue centers) are significantly flatter than faint ones, this is not the case for the nucleated dEs (again without disks or blue centers). Given the lower stellar population age of dE(nN)s \\citep{Rakos2004}, and their location in regions of lower density, a possible explanation would be that the brighter dE(nN)s still need to experience further dynamical heating through tidal interactions, before they become as round as the dE(N)s. For the faint dE(nN)s, though, one would have to assume that they were more susceptible to such interactions, and therefore already have a rounder shape. On the other hand, the correlation of the projected shape of dE(N)s and their heliocentric velocity \\citep{Lisker2009velo} includes both bright and faint objects, and seems to imply that the roundest shapes are achieved only when the galaxies' orbits are significantly circularized. To pursue the ``unification'' of dE subclasses, another requirement is that the dE(nN)s would need to form nuclei rather soon, before their stellar population ages are comparable to today's dE(N)s. (Or, alternatively, there might have been some mechanism in the past that supported nucleus formation, but is not as efficient anymore today.) Perhaps nuclei are currently being formed in the centers of the dEs with blue cores, where we are witnessing ongoing star formation \\citep{p2}. This would lead to nuclei with younger stellar populations than those of their host galaxies. However, \\citet{Cote2006} found dE nuclei to have intermediate to old ages, and \\citet{Lotz2004} measured similar colours of nuclei and dE globular clusters (GCs). Alternatively, most nuclei could form through coalescence of GCs, which would indeed be a more efficient process in the central cluster regions \\citep{Oh2000}. Still, some effect would have to explain why the ratio of dE(N)s and dE(nN)s increases strongly with increasing dE luminosity \\citep{san85b}, especially if the dE(nN)s were to be the immediate progenitors of the dE(N)s. Similar to the above argument about nucleus formation, the spatial distribution of dE(nN)s would soon have to become significantly more centrally concentrated.\\linebreak \\citet{Conselice2001} derived a two-body relaxation time for Virgo dEs of much more than a Hubble time. However, two-body relaxation might be too simple for the real situation. These issues are probably best investigated with cosmological simulations, by ``flagging'' different groups or individual galaxies when they enter a cluster, and following the evolution of their combined (observable) state over various timescales. Current and future generations of simulations\\linebreak \\citep[e.g.][]{BoylanKolchin2009} certainly provide such a possibility. Obviously, many ifs and thens are necessary to explain all dEs by a single formation channel, and several authors have argued in favour of multiple channels (\\citealt{p4}; \\citealt{Poggianti2001}; \\citealt{Rakos2004};\\linebreak \\citealt{vanZee2004b}). Nevertheless, several aspects of dEs still need to be explored both theoretically and observationally, like the presence and characteristics of their globular clusters and what they can tell us about their evolutionary history \\citep[see the discussion in][]{Boselli2008}. \\subsection{A plea for the study of early-type dwarfs} Galaxy mass is one of the main drivers of galaxy evolution and appearance. Internal physical processes like supernova feedback \\citep{DekelSilk1986} or the dynamical response to gas loss \\citep{YoshiiArimoto1987} have a stronger impact on dwarf galaxies than on giants, making dwarfs valuable test objects for galaxy formation models\\linebreak \\citep{Janz2008,Janz2009a}. Nevertheless, if we want to use them as probes of the physical mechanisms that govern galaxy evolution, we need to understand the origin(s) of this most abundant galaxy population of clusters --- which is a difficult task. The intriguing complexity of dEs is contrasted with the moderate amount of high-quality observational data. Now that the largest cosmological simulations and their semi-analytic models begin to reach significantly into the dwarf regime, it is essential to build complete observational samples providing a thorough characterization of fundamental scaling relations at low masses, involving structure, kinematics, and stellar population properties. These will provide indispensable benchmarks for new generations of galaxy models, as well as for future studies of dEs at significantly earlier epochs with extremely large telescopes." }, "0910/0910.4464_arXiv.txt": { "abstract": "{Cyclotron resonant scattering features are an essential tool for studying the magnetic field of neutron stars. The fundamental line provides a measure of the field strength, while the harmonic lines provide information about the structure and configuration of the magnetic field. Until now only a handful of sources are known to display more than one cyclotron line and only two of them have shown a series of harmonics.} {The aim of this work is to see the first harmonic cyclotron line, confirming the fundamental line at $\\sim$22 keV, thus increasing the number of sources with detected harmonic cyclotron lines.} {To investigate the presence of absorption or emission lines in the spectra, we have combined \\emph{RXTE} and \\emph{INTEGRAL} spectra. We modeled the 3--100 keV continuum emission with a power law with an exponential cut off and look for the second absorption feature.} {We show evidence of an unknown cyclotron line at $\\sim$47 keV (the first harmonic) in the phase-averaged X-ray spectra of 4U~1538$-$52. This line is detected by several telescopes at different epochs, even though the S/N of each individual spectrum is low.} {We conclude that the line-like absorption is a real feature, and the most straightforward interpretation is that it is the first harmonic, thus making 4U 1538$-$52 the fifth X-ray pulsar with more than one cyclotron line.} ", "introduction": "Cyclotron resonant scattering features (CRSFs), usually referred to as `cyclotron lines', have proved to be powerful tools for directly studying the magnetic field in neutron stars. CRSFs are present in the hard X-ray spectra of several X-ray pulsars and originate in the ``cyclotron process'' under extreme conditions. Through $E_{cyc}=11.6B_{12} \\times (1+z)^{-1}$ keV (the `12-B-12 law', where $z$ is the gravitational redshift), an energy of the fundamental feature in the hard X-rays indicates that the magnetic fields are rather strong ($ B \\sim 10^{12}$\\,G). Under such conditions, the interaction of the electrons and radiation must be treated quantum-mechanically. The behaviour of an electron in a strong magnetic field implies that the electron energy must be quantized in so-called Landau levels. These absorption features stem from the resonant scattering of photons by electrons, also referred to as cyclotron lines. While the fundamental energy of the cyclotron line provides valuable information about the magnitude of the field, it is only through the detection and the analysis of the harmonic lines that we can get direct information about the geometrical configuration of the B field (\\cite{harding} 1991; \\cite{araya} 2000; \\cite{schonherr07} 2007). However, to date, only in a handful of systems have harmonic lines been discovered, and only two systems have shown more than two (\\cite{santangelo99} 1999 \\& \\cite{coburn05} 2005). It is therefore paramount to add as many systems to this selected group as we can. In this work, we present a spectral analysis of the high mass X-ray binary pulsar 4U 1538$-$52. It is an eclipsing system consisting of the B0 I supergiant star QV Nor and a neutron star with an orbital period of $\\sim$3.728 days (Clark \\cite{clark}). The orbital eccentricity is $\\sim$0.08 (Corbet et al. \\cite{corbet}), although more recently a higher value of $\\sim$0.17 was deduced by Clark (\\cite{clark}). The X-ray eclipse lasts $\\sim$0.6 days (Becker et al. \\cite{becker}). The system is fairly bright in X-rays. The estimated flux is $\\sim (5-20)\\times 10^{-10}$ erg s$^{-1}$ cm$^{-2}$ in the 3$-$100 keV range (Rodes~\\cite{rodesPhD}). Thus, assuming a distance of the source of $\\sim$5.5 kpc (Becker et al. \\cite{becker}, Parkes et al. \\cite{parkes}) and an isotropic emission, the luminosity follows $\\sim (2-7)\\times 10^{36}$ erg/s. The magnetized neutron star has a spin period of $\\sim$529 s (Davison \\cite{davison}; Becker et al. \\cite{becker}). \\begin{figure*}[h!t] \\centering \\includegraphics[angle=-90,width=9cm]{12815fg1.ps} \\includegraphics[angle=-90,width=9cm]{12815fg2.ps} \\caption{Combined spectrum and model of data obtained with \\emph{PCA} (3$-$20 keV) and \\emph{HEXTE} (17$-$100 keV). Both data sets belong to the run carried out in 2001 and their orbital phases are 0.53 and 0.66, respectively. Bottom panels show the residuals in units of $\\sigma$ with respect to the model (see Section~\\ref{RXTEanalysis} for details). } \\label{rxte} \\end{figure*} The pulse-phase averaged X-ray spectrum of 4U 1538$-$52 has usually been described either by an absorbed power law modified by a high energy cutoff, a power law modified by a Fermi-Dirac cutoff, or by two power laws with indices of opposite sign multiplied by an exponential cutoff (the NPEX model, Mihara~\\cite{mihara}; Rodes et al. \\cite{rodes1}). In addition to these continuum models, an iron fluorescence line at $\\sim$6.4 keV and a cyclotron resonant scattering feature at $\\sim$20 keV discovered by \\emph{Ginga} (Clark et al. \\cite{clark90}) are needed to describe the data. The variability of this CRSF was studied by Rodes-Roca et al. (\\cite{rodes2}). Rossi X-ray Timing Explorer (\\emph{RXTE}) (Coburn \\cite{coburn}) and \\emph{BeppoSAX} data (Robba et al. \\cite{robba}) did not show evidence of the first harmonic at $\\sim$40 keV. Robba et al. (\\cite{robba}) presented some evidence of an absorption feature around 50 keV; however, because of the lack of a signal-to-noise ratio of the spectrum at these energies, the feature could not be confirmed. In this paper, we report on the 3--100 keV analysis based on the observations of 4U 1538$-$52 performed by the \\emph{RXTE} and \\emph{INTEGRAL} satellites. In Sect.~\\ref{data} we describe the observations and data analysis. In Sect.~\\ref{analyse} spectral analysis are presented, and summarized in Sect.~\\ref{conclusion}. ", "conclusions": "\\label{conclusion} We presented the spectral analysis of 4U 1538$-$52 using data from \\emph{RXTE} and \\emph{INTEGRAL}. We present evidence for a previously unknown absorption line like feature in the phase-averaged spectrum of the source. As we have shown in Section~\\ref{analyse}, we have been able to achieve a good fit to the phase-averaged spectra by including a Lorentzian absorption line at $\\sim$47 keV into the model (see Fig.~\\ref{ftest}). This absorption line is clearly visible whenever the signal-to-noise ratio in the spectrum is good enough to allow an analysis of the data. The most straightforward interpretation for this feature is that it is the first harmonic of the $\\sim$ 22\\,keV fundamental CRSF. According to the theory, cyclotron lines are due to the resonant scattering of photons by electrons whose energies are quantized into Landau levels by the magnetic field (M\\'esz\\'aros \\cite{meszaros}). The quantized energy levels of the electrons are harmonically spaced in the first order, such that the first harmonic line should be placed at twice the energy of the fundamental line, i.e. $ 2\\times E_{\\rm cyc}\\approx 42$ keV. In reality, however, the coupling factor between the fundamental and first harmonic is with $\\sim 2.20$ slightly higher than 2.0. This anharmonic spacing, however, has been observed already in several systems where more than one line is present. As explained by \\cite{schonherr07} (2007), the relativistic photon-electron scattering already produces some anharmonicity, because photons with energies close to the Landau levels may not escape the plasma if their energies are not changed by inelastic scattering. This, however, can not be the only reason because some systems show an anharmonic spacing larger than that predicted by this effect. A possibility to explain this difference is to take into account that the optical depths of the fundamental and the first harmonic could be different if they are formed at different heights above the neutron star. With increasing height, the strength of the magnetic field decreases resulting in a different CRSF energy. Another possibility is to consider a displacement of the magnetic dipole which would also explain the difference of energy of the two lines if the lines originate from the different poles of the neutron star. Therefore a significant phase dependence of the strength of the both lines is expected, however, the low signal-to-noise ratio at higher energies prevents us to test this hypothesis with the current data sets." }, "0910/0910.1846_arXiv.txt": { "abstract": "We present new advances in the spectral extraction of point-like sources adapted to the \\textit{Infrared Spectrograph} onboard the \\textit{Spitzer Space Telescope}. For the first time, we created a super-sampled point spread function of the low-resolution modules. We describe how to use the point spread function to perform optimal extraction of a single source and of multiple sources within the slit. We also examine the case of the optimal extraction of one or several sources with a complex background. The new algorithms are gathered in a plugin called \\texttt{AdOpt} which is part of the \\texttt{SMART} data analysis software. ", "introduction": "The ideal spectral extraction algorithm for point-like sources yields the maximum signal-to-noise ratio (S/N) while at the same time preserving the spectrophotometric fidelity. The standard extraction method is based on co-adding the flux in the cross-dispersion direction within a window large enough to contain (most of) the source flux. This method is known as ``tapered-column\" in the ``Spectroscopy Modeling Analysis and Reduction Tool\" (\\texttt{SMART}\\footnote{\\texttt{SMART} is available at \\textit{http://isc.astro.cornell.edu/smart/}.}, Higdon et al.\\ 2004) and is equivalent to the ``regular extraction\" of the \\textit{Spitzer} IRS Custom Extractor (\\texttt{SPICE}), provided by the \\textit{Spitzer Science Center} (\\textit{SSC}). Although this generally produces satisfactory results, the inclusion of noisy pixels which do not contain a significant fraction of the flux inevitably tends to degrade the quality of the extracted spectrum. This is because every pixel in the extraction window is given the same weight. Moreover, the extraction window is part of a pseudo-rectangle which is defined as a zone in the detector array where the wavelength is uniform (Figure\\,\\ref{fig:pseudo}). When a quadrilateral boundary crosses a pixel, the signal is assumed to be evenly distributed within that pixel, which can lead to artificial flux variations in the extracted spectrum. Because of the interplay of the angled spectral trace and the widening extraction aperture, this error oscillates with wavelength with an amplitude which decreases toward longer wavelengths. Knowledge of the point spread function (PSF) of the instrument can solve these problems, as it enables the so-called optimal extraction technique which weights the extracted data by the S/N of each pixel (Horne 1986). Optimal extraction therefore significantly reduces the statistical noise in the final spectrum compared to more typical extraction algorithms, which weight all the data within the window equally. So far, efforts on optimal extraction for the \\textit{IRS} have made use of template PSFs or analytical PSFs to fit the cross-dispersion profile of the data: \\begin{itemize} \\item Virtually any spectrophotometric standard star can be used to derive a template PSF, which then allows an empirical estimate of the PSF at the reference positions (referred to as ``\\textit{nod}\" positions\\footnote{See the \\textit{Spitzer/IRS} observer's manual at \\textit{http://ssc.spitzer.caltech.edu/documents/SOM/}}). This method is currently used by the software \\texttt{SPICE} (see Narron et al.\\ 2007). The resulting extraction undeniably provides better S/N as compared to a tapered column extraction, although the corresponding algorithm is sensitive to cross-dispersion offsets between a source position and the nod position. A simple shift of the PSF is not sufficient to acquire the best S/N possible, and in some cases, low-frequency oscillations can appear in the spectrum due to the misalignment. The reasons are inherent to the data used to compute the PSF template since the latter is created in a specific observational mode (default positions along the slit, default data sampling). \\item Analytical PSFs allow estimating the instrumental profile for any position along the slits. The independent effort from the ``Core to Disks\" legacy program (Evans et al.\\ 2003) uses such PSFs along with an extended emission background which is determined on-the-fly (see Lahuis et al.\\ 2007). Analytical PSFs can bear some uncertainties due to the lack of knowledge of the exact instrumental profile. \\end{itemize} \\begin{figure}[b!] \\epsscale{1.0} \\plotone{fig1.eps} \\caption{The standard extraction aperture for a single wavelength, superimposed on the grid defined by the pixels on the detector array. The quadrilateral represents the extraction window, which is a ``tapered-column\" whose width increases proportionally with wavelength to account for the varying point spread function. The tilt of the quadrilateral reflects the fact that the row axis of the detector array is not parallel to a line of constant wavelength.\\label{fig:pseudo}} \\end{figure} We have developed a new optimal extraction algorithm to be used with either of the low-resolution modules of the \\textit{Infrared Spectrograph} (\\textit{IRS}, Houck et al.\\ 2004) onboard the \\textit{Spitzer Space Telescope} (Werner et al.\\ 2004). The new extraction algorithm is available \\textit{via} the plugin \\texttt{AdOpt}\\footnote{The documentation is available at \\textit{http://isc.astro.cornell.edu/SmartDoc/SmartOptimal}.}, part of the new release of \\texttt{SMART}. In a nutshell, an empirical super-sampled PSF has been constructed for each row of the detector array and a multi-linear regression algorithm is used to weight the pixels and derive the flux from the source. The optimal extraction is based on detector rows rather than pseudo-rectangles to treat each pixel as indivisible and thus avoid uncertainties due to the lack of knowledge on the pixel response function. The algorithm presented in this paper produces a considerably higher S/N, up to a factor of $\\sim2$ for faint sources, than the current extraction method available in \\texttt{SMART} for point-like sources (``tapered column\"). An additional advantage of using a super-sampled PSF is that it remains valid anywhere along the aperture in the cross-dispersion direction. We found that the optimal extraction method is extremely sensitive to offsets between a source position and the nominal position. For this reason a new algorithm to locate the source in the slit has been implemented, with a precision of better than a twentieth of a pixel. It is also now possible to extract spectra of spatially blended sources, which is crucial when dealing with crowded regions, such as stellar clusters or nearby galaxies. Such as what can be achieved using iterative techniques (Lucy \\& Walsh 2003), the cross-dispersion profile of the data is decomposed into its components, including the spatial profiles of any number of sources in the slit and the extended background emission. Finally, it is also possible to extract sources with significant offset in the dispersion direction by calculating a modified PSF on-the-fly. In the following section the basic steps to construct the PSF are given. Section\\,\\ref{sec:method} details the mathematical description of the method and its application for extracting multiple sources are detailed. Section\\,\\ref{sec:applications} briefly explains some specific applications. ", "conclusions": "We have presented the optimal extraction algorithm adapted to the \\textit{Infrared Spectrograph} onboard \\textit{Spitzer}. Besides providing a significant increase in the signal-to-noise ratio as compared to the regular ``tapered\" extraction, the new algorithm allows extraction of multiple sources at any location within the aperture. The optimal extraction is also implemented for sources shifted in the dispersion direction. Finally, it is possible to perform an optimal extraction with a complex extended background emission. The \\texttt{AdOpt} plugin is released with the \\texttt{SMART} package (versions equal or later than 8.0). The code is available at the Infrared Science Center website, along with extensive documentation. \\texttt{SMART} and the optimal extraction will be maintained in the future, with a possible inclusion of the optimal extraction for the high-resolution modules." }, "0910/0910.1241_arXiv.txt": { "abstract": "Using new and published photometric observations of $\\mu^{1}~{\\mbox{Sco}}$~(HR~6247), spanning 70~years, a period of 1.4462700(5)~days was determined. It was found that the epoch of primary minimum suggested by Shobbrook at HJD~2449534.178 requires an adjustment to HJD~2449534.17700(9) to align all the available photometric datasets. Using the resulting combined-data light-curve, radial velocities derived from IUE data and the modelling software \\phoebe{}, a new system solution for this binary was obtained. It appears that the secondary is close to, or just filling, its Roche-lobe. ", "introduction": "$\\mu^{1}~{\\mbox{Sco}}$ (HR~6247; HD~151890; HIP~82514) was only the third spectroscopic binary to be discovered and is listed in the GCVS \\citep{Samus04} as an eclipsing binary variable. Over the intervening years there have been a number of detailed measurements and several significant studies. It is now classified as a semi-detached~(sd) binary as one component is believed to fill its Roche lobe. \\citet{CesterFedel77} considered $\\mu^{1}~{\\mbox{Sco}}$, to be unusual owing to: \\begin{enumerate} \\item their determination of an apparently high mass-ratio; \\item indications that the secondary component has a larger radius than the primary and overflows its Roche lobe; \\item both components appear to lie on the main sequence; \\item it being a member of the Scorpius-Centaurus cluster thus indicating an age of no more than $10^7$~years hence the possibility this system has just arrived on the main sequence. \\end{enumerate} A comprehensive analysis of $\\mu^{1}~{\\mbox{Sco}}$ was undertaken to determine the best estimate of period from all available photometric data, establish a suitable epoch and calculate the key parameters defining the system. The photometric data used included published measurements from: \\begin{enumerate} \\item \\citet{RudnickElvey38}, \\item \\citet{vanGent39}, \\item \\citet{Stibbs48}, \\item Tycho collected between~1990 and~1993 \\citep{ESA97,OchsenbeinBauer00}, \\item Hipparcos collected between~1990 and~1993 \\citep{ESA97,OchsenbeinBauer00}, \\item \\citet{Shobbrook04}, \\end{enumerate} and new measurements taken in 2006, 2007 and 2008 by one of the authors~(Moon). The photometric data was then combined with radial velocities obtained by \\citet{SticklandSahade96} from IUE spectral data~\\citep{IUE09} and the resulting dataset analysed using the software program \\phoebe{} \\citep{Prsa03,PrsaZwitter05c,PrsaGuinan08,Phoebe09} based on the established \\citet{WilsonDevinney71} theoretical construct \\citep{KallrathMilone99}. ", "conclusions": "The light variations of $\\mu^{1}~{\\mbox{Sco}}$ appear to be stable over the 70~years for which photometric data exists. Using the comprehensive dataset of more than 500~photoelectric measurements assembled here, the period was determined to be 1.4462700(5)~day. A suitable epoch based on the primary minimum is suggested to be HJD~2449534.17700(9). Using all readily available data, and the \\phoebe{} software package, the data for $\\mu^{1}~{\\mbox{Sco}}$ could be best represented by considering the system to be a semi-detached binary in which the size of the secondary is close to, or fills, its Roche Lobe. The system properties determined from comprehensive, iterative modelling using \\phoebe{} indicate that the masses of the components may be somewhat less than previously thought. With the secondary appearing to fill its Roche lobe it maybe possible to obtain spectroscopic evidence of gaseous streams in the $\\mu^{1}~{\\mbox{Sco}}$ system noting that Doppler tomography has produced indirect images of gas flows in interacting binaries \\citep{Richards06}. However, for Algol-type binaries the gaseous streams are faint relative to the main sequence primary star and their detection challenges current techniques. Improvements to techniques would likely be required to detect such gas flows in $\\mu^{1}~{\\mbox{Sco}}$. There are currently only limited radial velocity data available for the $\\mu^{1}~{\\mbox{Sco}}$. Given that the radial velocity data provides initial values for parameters such as the semi-major axis, masses and radii, further data could be useful in improving the modelling of this binary system. Additionally, photometric measurements near phases 0.1, 0.4, 0.6 and 0.9 could aid refinement of the current theoretical model through better defining some apparent features. With the semi-major axis of the orbit subtending 0.39~mas, this star may be a suitable target for the Sydney University Stellar Interferometer~(SUSI)." }, "0910/0910.3250_arXiv.txt": { "abstract": "An accurate geometric distortion solution for the \\textit{Hubble Space Telescope} UVIS-channel of Wide Field Camera 3 is the first step towards its use for high precision astrometry. In this work we present an average correction that enables a relative astrometric accuracy of $\\sim$1 mas (in each axis for well exposed stars) in three broad-band ultraviolet filters (F225W, F275W, and F336W). More data and a better understanding of the instrument are required to constrain the solution to a higher level of accuracy. ", "introduction": "\\label{sec1} The accuracy and the stability of the geometric distortion\\footnote{ A specification is needed. With the term ``geometric distortion'', or GD, which we will use hereafter, we are lumping together several effects under the same term: the optical field-angle distortion introduced by camera optics, light-path deviations caused by the filters, non-flat CCDs, alignment errors of CCDs on the focal plane, etc..} (GD) correction of an instrument is at the basis of its use for high precision astrometry. The particularly advantageous conditions of the {\\it Hubble Space Telescope (HST)} observatory make it ideal for imaging-astrometry of (faint) point sources. The point-spread functions (PSFs) are not only sharp and (essentially) close to the diffraction limit --which directly results in high precision positioning-- but also the observations are not plagued by atmospheric effects (such as differential refraction, image motion, differential chromatic refraction, etc), which severely limit ground-based astrometry. In addition to this, {\\it HST} observations do not suffer from gravity-induced flexures on the structures of the telescope (and camera), which add (relatively) large instabilities in the GD of ground-based images, and make its corrections more uncertain. Last May 14, the brand-new {\\it Wide Field Camera 3} (WFC3) was successfully installed during the {\\it Hubble Servicing Mission 4} (SM4, May 12-24 2009). After a period of intense testing, fine-tuning, and basic calibration, last September 9th, 2009, the first calibration- and science-demonstration images were finally made public. Our group is active in bringing {\\it HST} to the {\\it state of the art} of its astrometric capabilities, that we used for a number of scientific applications (e.g.\\ from King et al.\\ 1998, to Bedin et al.\\ 2009, first and last accepted papers). Now that the {\\it ``old''} ACS/WFC is successfully repaired, and that the new instruments are installed, our first step is to extend our astrometric tools to the new instruments (and to monitor the old ones). This paper is focused on the geometric distortion correction of the \\textit{UV/Optical} \\textit{(UVIS)} channel of the WFC3. Since the results of these efforts might have some immediate public utility (e.g.\\ relative astrometry in general, stacking of images, UV-identification of X-counterparts such as pulsars and CVs in globular clusters, etc.), we made our results available to the WFC3/UVIS user-community. We immediately focused our attention on a deep UV-survey of the core of the Galactic globular cluster $\\omega$~Centauri (NGC~5139), where some well dithered images were collected. The dense --and relatively flat-- stellar field makes the calibration particularly suitable for deriving and monitoring the GD on a relatively small spatial scale. In addition, while most of the efforts to derive a GD correction will be concentrated on relatively redder filters, we undertook a study to determine the GD solutions of the three bluest broad-band filters (with the exception of F218W): F225W, F275, and F336W. The WFC3/UVIS layout is almost indistinguishable from that of ACS/WFC\\footnote{\\sf http://www.stsci.edu/hst/wfc3}:\\ two E2V thinned, backside illuminated and UV optimized 2k$\\times$4k CCDs contiguous on the long side of the chip, and covering a field of view (FoV) of $\\sim160$$\\times$$160$ arcsec$^2$. The $\\omega$~Cen data set used here consists of 9$\\times$350~s exposures in each of the filters F225W, F275W, and F336W. The nine images follow a squared 3$\\times$3 dither-pattern with a step of about 40 arcsec (i.e.\\ $\\sim$1000 pixels), and were all collected on July 15, 2009. We downloaded the standard pipe-line reduced {\\sf FLT} files from the archive. The {\\sf FLT} images are de-biased and flat-field corrected, but {\\it no} pixel-resampling is performed on them. The {\\sf FLT} files are multi-extension fits (MEF) on which the first slot contains the image of what --hereafter-- we will call chip 1 (or simply [1]). The second chip, instead, is stored in the fourth slot of the MEF, and we will refer to it as chip 2 (or [2]). [Note that others might choose a different notation]. Our GD corrections refer to the raw pixel coordinates of these images. The fluxes and positions were obtained from a code mostly based on the software {\\sf img2xym\\_WFI} by Anderson et al.\\ (2006). This is essentially a spatially variable PSF-fitting method. We were pleased to see that for the WFC3/UVIS images of this data set the PSFs were only marginally undersampled. Left panel of Figure~\\ref{fig:0A} shows a preliminary color-magnitude diagram in the three filters for the bright stars in the WFC3/UVIS data set. In a future paper of this series we will discuss the PSF, its spatial variation and stability, as well as L-flats\\footnote{Residual low-frequency flat-field structure (L-flat) cannot be accurately determined from ground-based calibration data or internal lamp exposures. L-flats need to be determined from on-orbit science data, for example from multiple observations of stellar fields with different pointings and roll angles (van der Marel 2003).}, pixel-area corrections, and recipes for deep photometry in stacked-images. \\begin{figure*}[t!] \\centering \\includegraphics[width=14cm]{fig01r.ps} \\caption{ {\\it Left:} Preliminary color-magnitude diagram of the bright stars in the new WFC3/UVIS data set (fluxes are neither pixel-area- nor L-flat-corrected). {\\it Right:} Color-magnitude diagram of the stars in our ACS/WFC master frame (from Villanova et al.\\ 2007). Both plots are in instrumental magnitudes. } \\label{fig:0A} \\end{figure*} \\begin{figure*}[t!] \\centering \\includegraphics[width=14cm]{fig02.ps} \\caption{{\\it Left:} Color-magnitude diagrams of the high S/N stars in common with the ACS/WFC master frame (F435W), actually used to derive the geometric distortion correction, for each of the WFC3/UVIS filters (F225W, F275W, and F336W). } \\label{fig:0B} \\end{figure*} ~\\\\ \\newpage ~\\\\ \\newpage ~\\\\ \\newpage ", "conclusions": "\\label{sec:3} By using a limited (but best available) number of exposures with large dithers, and an existing ACS/WFC astrometric flat field, we have found a set of third-order correction coefficients to represent the geometric distortion of WFC3/UVIS in three broad-band ultraviolet filters. The solution was derived independently for each of its two CCDs. The use of these corrections removes the distortion over the entire area of each chip to an average accuracy of $\\sim$0.025 pixel (i.e.\\ $\\sim$1 mas), the largest systematics being located in the $\\sim$200 pixels closest to the boundaries of the detectors (and never exceeding 0.06 pixels). We advise the use of the inner parts of the detectors for high-precision astrometry. The limitation that has prevented us from removing the distortion at an even higher level of accuracy is the lack of enough observations collected at different roll-angles and dithers which could enable us to perform an auto-calibration. Nevertheless, the comparison of the mid-2002 ACS/WFC positions with the new WFC3 observations corrected with our astrometric solutions are good enough to clearly show the internal motions of $\\omega$~Centauri. These proved to be in perfect agreement with the most recent determinations. We also derived the average absolute scale of the detector with an accuracy limited by the uncertainties in the plate-scale variations induced by the velocity aberration of the telescope motion in the Earth-Sun system. For the future, more data with a longer time-baseline are needed to better characterize the GD stability of {\\it HST} WFC3/UVIS detectors in the medium and long term." }, "0910/0910.1580_arXiv.txt": { "abstract": "We present studies of the collapse of neutron stars that undergo a hadron-quark phase transition. A spherical Lagrangian hydrodynamic code has been written. As initial condition we take different neutron star configurations taking into account its density, energy density and pressure distribution. The phase transition is imposed at different evolution times. We have found that a significant amount of matter on the surface can be ejected while the remaining star rings in the fundamental and first pressure modes. ", "introduction": "Black holes ($BHs$) and neutron stars ($NSs$) are certainly two major potential sources of gravitational waves ($GWs$). Unlike $BHs$, whose gravitational waveforms are specified essentially by their masses and angular momenta, the characteristics of the gravitational emission from $NSs$ depend on the properties of the nuclear matter. When nuclear matter is compressed to a sufficiently high density, it turns into uniform three-flavor ($u$, $d$, and $s$) strange quark matter (SQM), since it is expected that SQM may be more stable than nuclear matter. The deconfined quark matter appears when the density is so high that the nucleons are ``touching'' each other. At this point, when the number density of nucleons $n\\sim O(1\\ {\\rm fm}^{-3})$, the quarks lose their correlation with individual nucleons and appear in a deconfined phase. Since the density required for this to happen is not much higher than nuclear matter density ($0.16\\ {\\rm fm}^{-3}$), the dense cores of neutron stars are the most likely places where the phase transition to quark matter may occur astrophysically. It should be noted that strange quarks (in a confined phase) could already exist in neutron stars with a hyperon core\\cite{marranghello}. In principle, the existence of a thin crust of normal matter is possible at the surface of a strange star. On the other hand, if SQM is metastable at zero pressure, so that it is relatively more stable than nuclear matter only because of the high pressure in the cores of neutron stars, then the final products of the phase transition would be hybrid stars, which consist of quark matter cores surrounded by normal matter outer parts. The above arguments were made in \\cite{lin} and those authors used polytropic equation of state to describe the complex nuclear matter structure. We proceed similar calculations with a realistic equation of state based on a field theoretical model. We also make use of a spherical Lagrangian hydrodynamical code instead of the Eulerian model used in \\cite{lin}. ", "conclusions": "\\indent We have obtained some important results which seems to be in complete agreement with those obtained by \\cite{lin}. The equation of state is shown in Fig.1 for three different parameter $\\alpha$ which leads to different hadron-quark phase transition pressure and density. Three stages of evolution are shown in Fig.2. The star radius shrinks and the central density is increasing during the collapse. In Fig.3 one can find the density evolution of each shell. The oscillation of the inner shells is clear in the final stage of evolution as it is the ejection of the external shells. The fourier transform is shown in Fig.4. This result shows a peak very close to the one shown by \\cite{lin}, at 3.1kHz, which shall correspond to the fundamental mode as well as other peaks, corresponding to the first pressure modes. It is also easy to find that most of the energy is released in the fundamental mode, as always claimed. In Fig.5 the radius and density evolution are plotted against time in this same figure. Finally, after a future detection, one can try to fit the nuclear matter parameters to obtain the frequency of the detection, remembering that, if the frequency falls on the 3.0-3.4kHz bandwidth, the Brazilian Schenberg antenna will be able to identify the direction of the source, as shown in \\cite{mara}. \\ack{We would like to thank CNPq and Fapesp for financial support.}" }, "0910/0910.2126_arXiv.txt": { "abstract": "The cosmic evolution of the metal content of the intergalactic medium puts stringent constraints on the properties of galactic outflows and on the nature of UV background. In this paper, we present a new measure of the redshift evolution of the mass density of \\CIV, $\\Omega_{\\rm CIV}$, in the interval $1.5 \\la z \\la 4$ based on a sample of more than 1500 \\CIV\\ lines with column densities $10^{12} \\la N($\\CIV$) \\la 10^{15}$ cm$^{-2}$. This sample more than doubles the absorption redshift path covered in the range $z<2.5$ by previous samples. The result shows a significant increase of $\\Omega_{\\rm CIV}$ towards the lower redshifts at variance with the previously pictured constant behaviour. ", "introduction": "The cosmological mass density of \\CIV, $\\Omega_{\\rm CIV}$, observed as a function of redshift is a fundamental quantity closely related to the metal enrichment of the intergalactic medium (IGM). Its apparent lack of evolution in the redshift interval $z\\simeq [1.5,5]$ \\citep{songaila01,pettini03,BSR03} is puzzling since both the physical conditions of the IGM and the properties of the ionizing background are thought to evolve between these epochs. Remarkable efforts have been spent in recent years to extend the measure of $\\Omega_{\\rm CIV}$ to redshift larger than 5 \\citep{ryanweber06,simcoe06} where a decrease of the star formation rate density is observed \\citep{bunker}. If $\\Omega_{\\rm CIV}$ is dominated by the metals produced in situ by the observed star forming galaxies, we would expect a decrease of its value at those redshifts. Vice versa, the value of $\\Omega_{\\rm CIV}$ could remain constant if it reflects the metallicity of a diffuse medium pre-enriched at very high redshift. It should be noted, however, that this is a simplified scenario since, as redshift increases, the observed \\CIV\\ absorptions likely trace gas in structures of decreasing over-density and also the ionizing spectrum evolves in shape and intensity. As a consequence, the behaviour of $\\Omega_{\\rm CIV}$ could be different from that of $\\Omega_{\\rm C}$ and of the mean IGM metallicity \\citep[see e.g.][]{schaye03}. The most recent measurements of \\CIV\\ absorptions in spectra of QSOs at $z\\sim6$ seem to indicate a downturn in the \\CIV\\ mass density at $z>5$ \\citep{becker09,ryanweber09}, though based only on 3 detected \\CIV\\ lines. At redshift $z\\la 4.5$, a fundamental measurement of $\\Omega_{\\rm CIV}$ has been carried out by \\citet[][S01]{songaila01}. However, the redshift interval $1.5 5$ leaving substantially unchanged the other values \\citep[][OD08]{opp_dave08}. An effect which has not been taken into account up to now is the evolution in the shape of the UV background due to the \\HeII\\ re-ionization process expected at redshift $\\sim 3$. Naively, due to the increase in the number of free hard photons (ionizing \\CIV\\ into \\CV) at the end of the re-ionization epoch, we would expect a trend opposite to what is observed: a decrease of the amount of \\CIV\\ after $z\\sim3$. However, the details of this process are still under study \\citep[e.g.][]{bolton09,hm09} and other factors should be accounted for properly: the decrease in the number density of QSOs going towards lower redshifts and the fact that at low $z$ the regions traced by the strong \\CIV\\ absorbers (driving the evolution of $\\Omega_{\\rm CIV}$) could be only mildly affected by the cosmic UV background. The observed raise of the cosmic \\CIV\\ mass density in the redshift range $1.5-2.5$ puts a strong constraints on the models describing the interplay between galaxies and their surrounding medium, suggesting that something is still missing in the physical implementation of galactic feedback." }, "0910/0910.5706_arXiv.txt": { "abstract": "We discuss current cosmological constraints on axions, as well as future sensitivities. Bounds on axion hot dark matter are discussed first, and subsequently we discuss both current and future sensitivity to models in which axions play the role as cold dark matter, but where the Peccei-Quinn symmetry is not restored during reheating. ", "introduction": " ", "conclusions": "" }, "0910/0910.2310_arXiv.txt": { "abstract": "We present an abundance analysis of the star Cernis~52 in whose spectrum we recently reported the napthalene cation in absorption at 6707.4 {\\AA}. This star is on a line of sight to the Perseus molecular complex. The analysis of high-resolution spectra using a $\\chi2$-minimization procedure and a grid of synthetic spectra provides the stellar parameters and the abundances of O, Mg, Si, S, Ca, and Fe. The stellar parameters of this star are found to be $T_{\\mathrm{eff}} = 8350 \\pm 200$ K, \\loggl $= 4.2 \\pm 0.4$ dex. We derived a metallicity of $\\mathrm{[Fe/H]} = -0.01 \\pm 0.15$. These stellar parameters are consistent with a star of $\\sim 2$~\\Msun in a pre-main-sequence evolutionary stage. The stellar spectrum is significantly veiled in the spectral range $\\lambda\\lambda5150-6730$~{\\AA} up to almost 55 per cent of the total flux at 5150~{\\AA} and decreasing towards longer wavelengths. Using Johnson-Cousins and 2MASS photometric data, we determine a distance to Cernis~52 of 231$^{+135}_{-85}$~pc considering the error bars of the stellar parameters. This determination places the star at a similar distance to the young cluster IC~348. This together with its radial velocity, $v_r=13.7\\pm1$ \\kmso, its proper motion and probable young age support Cernis~52 as a likely member of IC~348. We determine a rotational velocity of $v\\sin i=65 \\pm 5$~\\kms for this star. We confirm that the stellar resonance line of \\ion{Li}{1} at 6707.8~{\\AA} is unable to fit the broad feature at 6707.4~{\\AA}. This feature should have a interstellar origin and could possibly form in the dark cloud L1470 surrounding all the cluster IC~348 at about the same distance. ", "introduction": "In a recent study of the diffuse interstellar bands (DIBs) toward a region of anomalous microwave emission, Iglesias-Groth et al. (2008) have shown evidence for the presence of the naphthalene cation, ${\\rm C}_{10}{\\rm H}_8^+$, the simplest Policyclic Aromatic Hydrocarbon (PAH), in the line of sight towards the moderately reddened star Cernis~52 (BD+31$^o$ 640, $V=11.4$, $E(B-V)=0.9$, Cernis 1993). The PAHs, including the naphthalene cation, have been also detected in a cometary dust sample returned to Earth by Stardust and in interplanetary dust particles, possibly of cometary origin (see the review by Li 2009). Cernis~52 is an early type star, classified as A3V by Cernis (1993), has coordinates $\\alpha = 03^{\\rm h}43^{\\rm m}00.3^{\\rm s}$ and $\\delta = +31^\\circ58'26''$ (J2000.0), and is located at an angular separation of less than one degree from the very young stellar cluster IC~348 in the Perseus OB2 molecular complex. The photometric distance to star Cernis~52 is 240 pc (Cernis 1993), consistent with a location in the molecular complex OB2 where the microwave emitting cloud is also most likely located (see Watson et al. 2005). The uncertainties prevent to conclude whether the star is embedded in a cloud of this young star forming region or lay behind. Iglesias-Groth et al. (2008) claimed that the relative strength of several diffuse interstellar bands detected in the spectrum of Cernis~52 was consistent with the presence of a molecular cloud in the line of sight. The anomalous microwave emission detected in the direction to Perseus (Watson et al. 2005, 2006) could be associated to electric dipole radiation of fast spinning hydrogenated carbon-based molecules in a molecular cloud, a mechanism originally proposed by Draine and Lazarian (1998). It is important to establish whether the cloud responsible for the excess color of Cernis~52 hosts the carriers of both, the diffuse interstellar bands observed in the spectrum of this star and the anomalous microwave emission detected in its line of sight. Iglesias-Groth et al. (2008) also noted the near-coincidence in wavelength between the interstellar cation's feature and the stellar Li I resonance line. To gain insight into the contribution of the stellar line to the observed broad feature, we undertook a thorough analysis by spectrum synthesis of Cernis 52's spectrum with two principal goals - a general abundance analysis for the star and spectrum synthesis fits to the spectrum around 6707~{\\AA} in order assess the Li I line's contribution. Here we present a precise determination of the stellar parameters and a detailed chemical composition study of the star Cernis 52. This abundance analysis should provide indications of the likely Li abundance for the star, i.e., if the star is not chemically peculiar should have a lithium abundance no greater than the cosmic Li abundance of A(Li)\\footnote{A(Li)$=\\log [N({\\rm Li})/N({\\rm H})]+12$}\\,$=3.3.$ In this paper we will discuss with special attention the naphthalene cation's feature at 6707.4~{\\AA} in the spectrum of Cernis~52 and investigate the relationship of this star with the Perseus star forming region. ", "conclusions": "We have performed a detailed chemical analysis of the star Cernis~52. We apply a technique that provides a determination of the stellar parameters, taking into account any possible source of veiling. We find $T_{\\mathrm{eff}} = 8350 \\pm 200$ K, $\\log (g/{\\rm cm~s}^2) = 4.2 \\pm 0.4$, $\\mathrm{[Fe/H]} = -0.01 \\pm 0.15$, and a veiling (defined as $F_{\\rm veil}/F_{\\rm total}$) of less than 55\\% at 5000 {\\AA} and decreasing toward longer wavelengths. The spectrum of Cernis~52 shows many features very likely related with the interstellar medium. In addition, we discover in photometric images a companion 1.7~mag fainter star at a distance of $0.818\\arcsec\\pm0.007\\arcsec$, but this star only contributes with 10\\% of the stellar flux in our spectra and hardly affect the stellar features of Cernis~52. The derived chemical abundances are roughly solar within their error bars. This prevent the star from being a chemically peculiar star. We have determined the radial velocity of Cernis~52 at $v_r=+13.7 \\pm 1$~\\kmso, being almost equal to the mean radial velocity of the young cluster IC~348. We have compared the stellar parameters with pre-main-sequence evolutionary tracks of solar metallicity and see that the star is consistent with being a pre-main-sequence A-type star with an age of $3-20$~Myr. We have estimated the distance to Cernis~52 using the available Johnson-Cousins and 2MASS photometric data, at 231$^{+135}_{-85}$~pc according to the stellar parameters and its error bars. This value also agrees with the distance to the cluster IC~348. The proper motion of Cernis~52, $(\\mu_\\alpha \\cos \\delta$, $\\mu_\\delta=(+7.22$,$-8.62)\\pm(1.5$,$1.6)$ mas~yr$^{-1}$, is consistent with the proper motion of IC~348. All these measurements make it likely that the star Cernis~52 (\\mbox{BD+31$^o$ 640}) belongs to the young cluster IC~348. We confirm that the feature at 6707.4~{\\AA} is not related with a stellar lithium line because the rotational velocity of the star, $v\\sin i=65 \\pm 5$~\\kmso, is too low to explain the broad feature associated with the naphthalene cation. Furthermore, the presence of a companion star cannot either explain this feature even for an abnormally high Li abundance, $\\log [N({\\rm Li})/N({\\rm H})]+12 > 5$ dex. As already stated in Cernis (1993), the interstellar features, in particular, the naphthalene cation, that appear in the spectrum of Cernis~52 may form in the dark cloud L1470 which covers all the cluster IC~348 and is at about the same distance." }, "0910/0910.5476_arXiv.txt": { "abstract": "We present {\\it Spitzer} observations of a sample of 12 starless cores selected to have prominent 24~\\micron\\ shadows. The {\\it Spitzer} images show 8 and 24~\\micron\\ shadows and in some cases 70~\\micron\\ shadows; these spatially resolved absorption features trace the densest regions of the cores. We have carried out a $^{12}$CO (2-1) and $^{13}$CO (2-1) mapping survey of these cores with the Heinrich Hertz Telescope (HHT). We use the shadow features to derive optical depth maps. We derive molecular masses for the cores and the surrounding environment; we find that the 24~\\micron\\ shadow masses are always greater than or equal to the molecular masses derived in the same region, a discrepancy likely caused by CO freeze--out onto dust grains. We combine this sample with two additional cores that we studied previously to bring the total sample to 14 cores. Using a simple Jeans mass criterion we find that $\\sim\\!2/3$ of the cores selected to have prominent 24~\\micron\\ shadows are collapsing or near collapse, a result that is supported by millimeter line observations. Of this subset at least half have indications of 70~\\micron\\ shadows. All cores observed to produce absorption features at 70~\\micron\\ are close to collapse. We conclude that 24~\\micron\\ shadows, and even more so the 70~\\micron\\ ones, are useful markers of cloud cores that are approaching collapse. ", "introduction": "Stars are born in cold cloud cores \\citep[e.g.,][]{bergin07}, where gas and dust are compressed to high enough densities to cause the core to start collapsing. One of the most critical questions in astronomy is what the physical conditions for core formation and collapse are. Therefore, observations of cores in the rare stage during, or just prior to, collapse are critical to constrain theories of the very earliest stages of star-formation \\citep[e.g.,][]{shu87,ballesteros03,myers05}. The initial process of collapse, or the transition from a stable core to a core with an embedded protostar, happens rapidly \\citep{hayashi66}; additionally, the cores are necessarily dense and cold. These conditions pose significant observational challenges to identify cores close to collapse. Because of the temperature ranges involved, $\\sim\\!10$~K \\citep[e.g.,][]{lemme96,caselli99,hotzel02}, and high densities, $\\sim$10$^5$ cm$^{-3}$ \\citep[e.g.,][]{bacmann00}, cold cloud cores can best be observed at far-infrared, submillimeter, and millimeter wavelengths \\citep[e.g.,][]{stutz09}. Such observations of dense cores show that most of them are close to equilibrium and not collapsing \\citep[e.g.,][]{lada08}. Furthermore, if the cores are not supported by thermal and turbulent pressure alone, then a modest magnetic field can halt collapse \\citep[e.g.,][]{kandori05,stutz07,alves08}. Millimeter line observations are the traditional way to search for collapsing cores \\citep[e.g.,][]{walker86}; however, there are ambiguities in the interpretation of such measurements \\citep[e.g.,][]{walker88,menten87,mundy90,narayanan98}. Here we use mid-infrared shadows as an alternative observational approach to the study of starless cores, one that does not depend on the interpretation of millimeter line profiles and how these profiles relate to the underlying velocity field of the core material. Additionally, our method is sensitive to higher column densities than traditional studies of starless cores by background star extinction, usually limited to A$_V$ $\\le$ 30 magnitudes \\citep[e.g.,][]{alves01,kandori05}. Cores that are isolated from regions of massive star formation are very useful to understand the initial stages of collapse into stars. These cores are free from complicating factors, e.g., massive core fragmentation, feed-back from high-mass stars, and protostellar outflows \\citep[e.g.,][]{lada03,dewit05,fallscheer09}, to name a few, that can have large effects on a study aimed at understanding how individual low mass stars form. We present a sample of 12 relatively isolated cores that were selected from {\\it Spitzer} MIPS observations of Bok globules and other star-forming regions to have prominent 24 $\\mu$m shadows, spatially resolved absorption features caused by the dense core material viewed in absorption against the interstellar radiation field. We also include 2 previously studied cores, CB190 \\citep{stutz07} and L429 \\citep{stutz09}, for a final sample of 14 cores. Such objects with 24 $\\mu$m shadows always show a counterpart 8 $\\mu$m shadow, analogous to the IR absorption features produced by more distant and massive structures termed infrared dark clouds \\citep[IRDCs; e.g.,][]{perault96,butler09,peretto09,lee09} and the ISO 7 $\\mu$m absorption features studied by \\citet{bacmann00}. We also present Heinrich Hertz Telescope (HHT) $^{12}$CO (2-1) and $^{13}$CO (2-1) on-the-fly (OTF) maps of regions $\\sim$ 10$'$ on a side surrounding each core. Our search method is efficient at identifying cores that are near collapse. In \\S~2 we describe the observations and data processing; in \\S~3 we derive optical depth maps from the 8 and 24 $\\mu$m images; in \\S~4 we present our mass measurements and we describe the stability analysis that we apply to the sample of cores; and finally, in \\S 5 we summarize our main conclusions. ", "conclusions": "We study the 8~\\micron, 24~\\micron, and 70~\\micron\\ shadows cast by a sample of 14 starless cores. We derive 24~\\micron\\ core masses and sizes; we apply a Jeans mass criterion, and attempt to account for turbulent and magnetic support in the cores, in order to assess the collapse state of each core. We caution that distance uncertainties can have a large effect on our Jeans mass analysis. In addition, we have obtained $^{12}$CO (2-1) and $^{13}$CO (2-1) OTF maps of the cores; the molecular core masses we derive are always less than the 24~\\micron\\ masses. This discrepancy is likely caused by freezeout onto dust grains. From this work we conclude that: \\noindent 1. 70\\% (10/14) of cores selected to have prominent 24~\\micron\\ shadows seem to be approaching collapse, indicating that this criterion selects dense, evolved cores. \\noindent 2. 50\\% (5/10) of cores that are classified as approaching collapse have indications of 70~\\micron\\ shadows; the 70~\\micron\\ data quality does not allow for a rigorous analysis of these shadows. \\noindent 3. All cores with indications of 70~\\micron\\ shadows have millimeter line profiles showing blue asymmetries, indicating that these long--wavelength shadow features are produced by very evolved cores. These cores are all likely to be close to collapse. \\noindent 4. Shadows at 24~\\micron\\ and especially at 70~\\micron\\ appear to be effective indicators of cores that are approaching collapse." }, "0910/0910.0854_arXiv.txt": { "abstract": "The Minimal Supersymmetric Standard Model has several flat directions, which can naturally be excited during inflation. If they have a slow (perturbative) decay, they may affect the thermalization of the inflaton decay products. In the present paper, we consider the system of udd and QLd flat directions, which breaks the $U(1)\\times SU(2)\\times SU(3)$ symmetry completely. In the unitary gauge and assuming a general soft breaking mass configuration, we show that for a range of parameters, the background condensate of flat directions can undergo a fast non-perturbative decay, due to non-adiabatic evolution of the eigenstates. We find that both the background evolution and part of the decay can be described accurately by previously studied gauged toy models of flat direction decay. ", "introduction": "\\label{sec:intro} Flat directions are generic features of supersymmetric theories. They are directions in field space along which the renormalizable part of the scalar potential vanishes. The Minimal Supersymmetric Standard Model (MSSM) and its extensions have a plethora of D and F-flat directions \\cite{Gherghetta:1995dv}, which are lifted due to supersymmetry breaking. During inflation, if their effective mass is small compared to the Hubble rate, the fields can develop large vacuum expectation values (VEV) along the flat directions of the potential \\cite{developvev,Ellis:1987rw}. This growth is bounded above by the non-renormalizable term with lowest dimension, which has the form $\\phi^d/M^{d-3}$ with $d \\geq 4$. If inflation is long enough, the growth will proceed up to $\\langle \\phi \\rangle \\sim (m_\\phi M^{d-3})^{1/(d-2)}$ \\cite{Dine:1995kz}. If all non-renormalizable terms allowed by gauge invariance are present, each class of flat direction will be lifted by the term with the smallest $d$ allowed by MSSM symmetries \\cite{Gherghetta:1995dv}. On the other hand, discrete symmetries may forbid some of such terms and a non-renormalizable term with a higher $d$ may determine the VEV of the flat direction. All possible flat directions in MSSM are excited by terms with $d \\le 9$ \\footnote{The term at which all flat directions are lifted may be different for the extensions of MSSM. For example, in $\\nu$MSSM, no flat direction survives beyond $d=6$ \\cite{Basboll:2009tz}.}. The formed condensate can have several cosmological implications: In the presence of phase dependent potential terms, the flat directions may source a finite baryon number density through the Affleck-Dine mechanism \\cite{Affleck:1984fy, Linde:1985gh, Allahverdi:2000zd}. It has also been suggested that they may be responsible for inflation \\cite{Allahverdi:2006iq}. Additionally, due to the large VEV of the flat directions, all the fields coupled to them acquire a large effective mass, slowing down the decays they mediate and resulting in a small perturbative decay rate. Typically, the perturbative decay of flat direction concludes after $\\sim 10^{11}$ rotations \\cite{Olive:2006uw}. These long lived flat directions also keep the gauge fields of broken symmetries (assumed to be all the Standard Model ones) heavy, suppressing the scatterings among the inflaton decay products, thus delaying their thermalization \\cite{Allahverdi:2005mz}. In addition, the energy density of the (relativistic) inflaton decay products may become sub-dominant over that of (massive) flat directions. The subsequent radiation stage will then be dominated by the thermal distribution of flat direction decay products, rather than those of the inflaton. These effects on thermalization require sufficiently large initial flat direction VEVs, which can be acquired only if non-renormalizable superpotential terms up to $d=11$ are absent \\cite{Olive:2006uw}. However, if the decay of the flat directions is controlled by non-perturbative effects, the effect on thermalization will be very different than the above picture. This possibility was first discussed in \\cite{Allahverdi:1999je}, in the framework of a toy model based on F-term type interactions. For this model, the frequencies of the particles coupled to the flat directions evolve adiabatically, not allowing a resonant decay. On the other hand, it was shown in \\cite{Olive:2006uw}, that the D-term potential provides non-trivial interactions among the perturbations through a non-diagonal and time dependent mass matrix. Even if the eigenvalues of this matrix evolve adiabatically, the diagonalization procedure itself may be non-adiabatic, due to a fast rotation of the eigenvectors. The resulting exponential decay of the condensate has a much higher rate than the perturbative one, giving a decay after $\\mathcal{O}(10)$ rotations of the flat directions. In \\cite{Olive:2006uw}, it was also argued that at least two or more flat directions need to be excited for this effect to be realized. The argument is as follows: Since the resonant effect occurs in the D-terms, only the perturbations coupled to the VEVs through the symmetry generators are counted. Out of these degrees of freedom, two per broken symmetry will correspond to a Higgs and a Goldstone. Furthermore, two more (light) degrees of freedom will decouple, corresponding to the real and imaginary parts of fluctuations along each flat direction. In order to have a non-adiabatic mixing, one needs additional light degrees of freedom that the condensate can decay into. To formulate, the number of remaining degrees of freedom present in the system will be \\begin{equation} \\left( \\begin{array}{c} {\\rm remaining}\\\\ {\\rm degrees} \\end{array} \\right) = \\left( \\begin{array}{c} {\\rm d.o.f.} \\\\ {\\rm in~D~terms} \\end{array} \\right) - 2\\,\\times \\left( \\begin{array}{c} {\\rm broken} \\\\ {\\rm symmetries} \\end{array} \\right)-2\\,\\times \\left( \\begin{array}{c} {\\rm flat} \\\\ {\\rm directions} \\end{array} \\right)\\,. \\label{counting} \\end{equation} As long as this number is zero, there will not be any room for non-perturbative decay. For instance, for the typical cases of single flat directions, no residual degree of freedom is present \\cite{Olive:2006uw}. On the other hand, there exist flat directions that are non-exclusive, i.e. they do not give a large mass to each other due to their VEVs. If the conditions to excite a single flat direction are present, one can expect that the whole set of flat directions non-mutually exclusive with that one is excited. If realized, such a case would provide the extra degrees of freedom into which the condensate may decay non-perturbatively. The longevity of single flat directions was later reiterated by the authors of \\cite{Allahverdi:2006xh}, where it was also argued that for the non-perturbative decay of multiple flat directions, one needs some degree of tuning of the initial VEVs: Since different flat directions may be lifted by different non-renormalizable terms in the superpotential \\cite{Gherghetta:1995dv}, one may in general expect hierarchical VEVs. Such a case reduces to a single flat direction, which decays only perturbatively. The maximum amount of hierarchy that can provide a non-perturbative decay depends on the ellipticity of the orbits of the VEVs in their complex plane. In later works, gauged toy models with two flat directions \\cite{Basboll:2007vt,Gumrukcuoglu:2008fk} and examples from MSSM \\cite{Basboll:2008gc} were studied, each verifying that multiple flat directions may decay non-perturbatively. Additionally, in \\cite{Gumrukcuoglu:2008fk}, the fast decay was shown to be realized for a range of VEV ratios of three orders of magnitude. This range was found to be a consequence of the phase dependent terms introduced in the fashion of \\cite{Affleck:1984fy}. However, the question of whether the toy models provide a good description of MSSM flat directions needs to be answered. For example, in \\cite{Gumrukcuoglu:2008fk}, the gauged toy models of four fields with only $U(1)$ or $SU(2)$ charges have been studied, yet in MSSM, no such flat direction configuration is possible and generically, for multiple flat directions, the field content has charges of all symmetries. Furthermore, the production of the remaining degrees of freedom (\\ref{counting}) may be suppressed if they acquire large masses through the F-terms. Therefore, the main goal of the present work is to find a concrete example from MSSM for which, the decay of the flat directions proceed analogously to the gauged toy model case. The flat directions in the models of \\cite{Gumrukcuoglu:2008fk} are decoupled at the background level, and each of the two VEVs evolve independently like two single flat directions. However, having independently evolving VEVs is not a requirement for non-perturbative decay. For instance, flat directions with coupled VEVs also have the necessary ingredients for decay \\cite{Basboll:2007vt, Basboll:2008gc}. The latter systems are much more complicated than the former ones and a precise answer requires extended numerical calculation. Our primary focus will be on the system of $u^cu^cd^c$ and $QLd^c$ flat directions and we will show that their decay can be described by the four field toy model of \\cite{Gumrukcuoglu:2008fk}. Additionally, we will address some issues arising from the assumption that the fluctuations along the flat directions are decoupled from the other modes. For instance, for the $QLd^c+LLe^c$ system, Ref. \\cite{Basboll:2008gc} claimed that the Higgses and the flat direction perturbations have non-adiabatic mixings. On the other hand, for the toy models of \\cite{Gumrukcuoglu:2008fk}, it was shown that the these degrees of freedom indeed decouple from the rest of the action. However, the latter result is a consequence of the assumption that the fields in a given flat direction have equal masses, an assumption not generically applicable to MSSM fields. If the fields have distinct masses, the flat direction perturbations are no longer decoupled from the Higgses. If these mixings are non-adiabatic, they may result in a non-perturbative decay, even if the counting (\\ref{counting}) leaves no extra degrees of freedom. For the models we consider, we show that these mixings have negligible effect and the flat direction perturbations decouple as described in \\cite{Olive:2006uw, Gumrukcuoglu:2008fk}. We will first generalize the single flat direction toy model of \\cite{Gumrukcuoglu:2008fk} to have arbitrary masses and verify that the flat direction does not decay non-perturbatively. The approximations and methods we adopt in this simple example will provide us the necessary tools for the background evolution of $u^c u^c d^c$ and $Q L d^c$ flat directions, which will also be studied with generic mass terms. The paper is organized as follows. In Section \\ref{sec:mssm}, we discuss different classes of multiple flat directions in MSSM over some examples which allow for non-perturbative decay, and determine which example is most likely to be described by the gauged $4$-field toy model. In Section \\ref{sec:formalism}, we review the formalism for the calculations of this decay. In Section \\ref{sec:1fd}, we generalize the single flat direction toy model with two complex scalar fields and $U(1)$ gauge field, to include arbitrary soft masses. This provides a basis for non-degenerate mass calculation of a more complicated model, carried out in the next section. In Section \\ref{sec:2fd} we present a complete study of $u^cd^cd^c+Q L d^c$ flat directions in MSSM, with arbitrary soft masses, where we compare the final results to the ones for the $4$-field toy model in \\cite{Gumrukcuoglu:2008fk}. The results are summarized and discussed in Section \\ref{sec:disc}. Finally, we include the technical steps of the calculation in the appendices at the end. ", "conclusions": "\\label{sec:disc} The primary aim of this paper was to see if the extensive numerical study carried out in the framework of $4$-field gauged toy model \\cite{Gumrukcuoglu:2008fk} can accurately describe a realistic case. We have shown that this toy model, which has only $U(1)$ symmetry, two mass parameters and no F-terms, provides a very good description of an example from MSSM, where both $u^cd^cd^c$ and $Q L d^c$ flat directions are present simultaneously. Specifically, both cases contain ``independent flat directions'', that is, the fields in one flat direction are decoupled from the others at the background level. As a result, the background equations of motion for the two models have the same form. Furthermore, we found that out of the $64$ real degrees of freedom in the problem, only the two decoupled actions $S_{s_{\\bar{1}},s_2, d_{\\bar{2}}}$ and $S_{b_{\\bar{1}},s_3, d_{\\bar{3}}}$, each consisting of 8 degrees, exhibit the non-adiabatic evolution of the eigenstates. At the limit of degenerate masses, that is, when the masses of the fields in a flat direction are identical, we found that $S_{b_{\\bar{1}},s_3, d_{\\bar{3}}}$ can be decomposed into two copies of the coupled system in the $4$ field gauged toy model. Therefore, it is safe to state that, in the degenerate mass case, the numerical results of the toy model provides an exact description of this part of the action. If the action $S_{s_{\\bar{1}},s_2, d_{\\bar{2}}}$ contributes to the non-perturbative decay as we expect, the resulting quanta will be different ones and their production will not effect the occupation numbers of generated $b_{\\bar{1}}$, $s_3$ and $d_{\\bar{3}}$ perturbations, as long as the linearized approximation holds. Another focus of the present paper was to understand the effect of different soft supersymmetry breaking masses to the non-perturbative production. For an exactly D-flat direction, the fields in that direction have the same soft mass. Conversely, if all the fields have different masses, the D-term will be non-zero and proportional to the fourth power of mass differences. As a reference calculation, we showed that for a single flat direction, the flat direction perturbations still decouple at the leading order, thus verifying that there is no non-perturbative decay in this case. The effect of different masses in the $u^cd^cd^c + Q L d^c$ example had similar consequences for the flat direction perturbations. Furthermore, the four field toy model with degenerate masses still describes the subsystem $S_{b_{\\bar{1}},s_3, d_{\\bar{3}}}$ exactly when the condition (\\ref{deltamcom}) holds. Therefore, based on the results of \\cite{Gumrukcuoglu:2008fk}, we conclude that the flat directions $u^cd^cd^c + Q L d^c$ decay in $\\mathcal{O}(10)$ rotations, also for this case. One issue about the example considered here is the hierarchy between the VEVs corresponding to each flat direction. The first non-renormalizable operator present for $u^cd^cd^c$ flat direction has $d=6$, whereas for $Q L d^c$, it has $d=4$ \\cite{Gherghetta:1995dv}, resulting in an initial VEV ratio of $\\sim10^3$. This ratio, although large, may still give rise to a production if the orbits have enough ellipticity and the mass ratio is large enough. For instance, in the numerical results of \\cite{Gumrukcuoglu:2008fk}, it was shown that for two flat directions with mass ratio $\\tilde{m}/m =7.63$, they decay within $20$ rotations for initial VEV ratio of $10^3$. If the orbits are closer to the radial one, one might still have a non-perturbative decay with a smaller mass ratio for the same VEV hierarchy. On the other hand, the non-renormalizable terms are model dependent, and they may be forbidden by some discrete symmetries. In fact, to recover the conditions of delayed thermalization \\cite{Allahverdi:2005mz}, one needs $\\phi_{\\rm in} \\gtrsim 10^{-2} M_p$, which requires all terms with $d < 11$ to vanish \\cite{Olive:2006uw}. Although the numerical results of \\cite{Gumrukcuoglu:2008fk} shows that the flat directions may decay non-perturbatively, it is still not clear how this effect changes the picture of thermalization. Specifically, the new quanta are produced in a resonant band with momenta $k \\lesssim m$ which is still non-relativistic. Since the variances are large, the gauge fields will still have large enough mass contributions to suppress the scatterings between inflaton decay products. On the other hand, tracking the evolution of the produced particle distribution is beyond the reach of the linearized calculation. One needs to control the back reaction effects to determine how fast the variances decrease. In such a computation, the distribution is likely to thermalize, possibly not in $\\mathcal{O}(10)$ rotations, but we expect it to be much earlier than $10^{11}$ rotations that is required by the perturbative decay \\footnote{Indeed, our preliminary results from lattice simulations of non-gauged toy model \\cite{Olive:2006uw} up to $\\mathcal{O}(100)$ rotations show the beginning of thermalization of the produced particles, i.e. the initial distribution with momenta $k \\lesssim m$ starts to extend toward high momentum region.}. It is clear that the linear study done in the present paper, along with \\cite{Olive:2006uw, Basboll:2007vt, Basboll:2008gc, Gumrukcuoglu:2008fk} are limited to the stages of the evolution until the production is significant, and they only provide a glimpse to the beginning of the non-perturbative decay. For instance, we found that the $u^cd^cd^c + Q L d^c$ example should result in a decay within $\\mathcal{O}(10)$ rotations for the range of parameters in \\cite{Gumrukcuoglu:2008fk}. However, once non-linear effects become important, the particles produced through different decoupled actions may interact and change the outcome of this study. Therefore, one should be cautious to extrapolate a non-linear study based on the toy model to a realistic one. The logical step to be taken next is to include higher order terms and study the effect on the decay time and thermalization of the produced quanta. In a recent study \\cite{Dufaux:2009wn}, the non-gauged toy model of \\cite{Olive:2006uw} was evolved on a lattice using the ClusterEasy code \\cite{Felder:2000hq}, verifying that the non-perturbative decay is still realized within the first few rotations. Another interesting result of \\cite{Dufaux:2009wn} was the calculation of the gravity waves, sourced by the quick decay of the flat directions. Their spectrum was found to fall naturally into Hz-kHz range and depending on the initial VEV of the flat directions, may potentially be within the reach of upcoming experiments, such as Advanced LIGO. For more realistic models with gauge fields, the resulting spectrum may be different, but since the mass scale of the flat directions is of order TeV, it will have a frequency range similar to that in the toy model. In a future work, we will address the effect of back-reaction and gravity wave production in the framework of gauged models. However, there are still some problems left that can be dealt with analytical tools. In \\cite{CyrRacine:2009qk}, it was shown that the non-perturbative decay of flat directions may have an observable effect through the amplification of curvature perturbations. In the context of natural supergravity inflation \\cite{Kawasaki:2000ws}, Ref. \\cite{Kaminska:2009wh} showed that the non-perturbative decay of flat directions allows the inflaton preheating to be realized in these models. On the other hand, even at the linearized level, we do not have a complete numerical study of the decay for ``overlapping flat directions'', although we have approximate calculations showing the non-adiabatic rotation occurs \\cite{Basboll:2007vt, Basboll:2008gc}. Specifically, the toy model of \\cite{Basboll:2007vt} is qualitatively different than the ones considered in the present work, as well as \\cite{Gumrukcuoglu:2008fk}, in the sense that the flat directions are coupled at the background level. A numerical analysis in the fashion of \\cite{Gumrukcuoglu:2008fk} would be very useful in understanding the time scale of the decay and the range of hierarchy that allows production. There are many examples with ``overlapping flat directions'' in MSSM which provide the necessary ingredients for non-adiabatic evolution (e.g. \\cite{Basboll:2008gc}), and it is an interesting challenge to find simple toy models that correctly describe at least parts of these examples." }, "0910/0910.2851_arXiv.txt": { "abstract": "name{Samenvatting}% \\def\\bibname{Bibliografie}\\def\\chaptername{Hoofdstuk}% \\def\\appendixname{Bijlage}\\def\\contentsname{Inhoudsopgave}% \\def\\listfigurename{Lijst van figuren}\\def\\listtablename{Lijst van tabellen}% \\def\\indexname{Index}\\def\\figurename{Figuur}\\def\\tablename{Tabel}% \\def\\partname{Deel}\\def\\enclname{Bijlage(n)}\\def\\ccname{Ter attentie van}% \\def\\headtoname{Aan}\\def\\headpagename{Pagina}% \\def\\today{\\number\\day\\space\\ifcase\\month\\or januari\\or februari\\or maart\\or% april\\or mei\\or juni\\or juli\\or augustus\\or september\\or oktober\\or% november\\or december\\fi \\space\\number\\year}% \\typeout{ >>>>> use hlatex209 for Dutch hyphenation <<<<< }} \\hyphenation{Schrij-ver Krij-ger Kuij-pers Bal-le-gooij-en} \\def\\warningoverprint #1{\\special{!userdict begin /bop-hook{gsave 100 600 translate -45 rotate /Times-Roman findfont 100 scalefont setfont 0 0 moveto 0.9 setgray (#1) show grestore}def end}} \\newcounter{onefig} \\newcounter{fignumber} \\newcount\\nocaptions \\newcount\\nofigures \\newcount\\figwidth \\newcount\\viewgraphs \\def\\figsfile{} \\def\\figspath{} \\def\\paper{} \\def\\format{} \\def\\figlabel{} \\long\\def\\nextfig#1{\\setcounter{figure}{\\value{fignumber}} \\addtocounter{fignumber}{1} \\ifnum \\viewgraphs=1 \\newpage \\pagestyle{empty} \\fi \\ifnum\\value{onefig}=0 #1 \\fi \\ifnum\\value{onefig}=\\value{fignumber} #1 \\fi} \\def\\figwidths#1#2{\\ifnum \\nocaptions=1 #2mm \\else #1mm \\fi} \\def\\paper#1{} % \\long\\def\\plotfig#1#2{\\ifnum \\nofigures=1 \\else #2 \\fi} \\long\\def\\captiontext#1{\\ifnum \\nofigures=1 \\raggedright \\fi \\ifnum \\nocaptions=1 \\paper \\ifnum \\viewgraphs=0 \\newline \\mbox{}\\hrulefill\\mbox{} \\newline \\newline label:~\\{\\figlabel\\} \\fi \\else \\ifnum \\nofigures=0 \\fi #1 \\fi} \\newcount\\panelwidth \\newcount\\panelheight \\newcount\\bxmin \\newcount\\bymin \\newcount\\bxmax \\newcount\\bymax \\newcount\\tbxmin \\newcount\\tbymin \\newcount\\tpanelwidth \\newcount\\tpanelheight \\newcount\\tpdif \\panelwidth=70 \\panelheight=70 % \\def\\panelsize #1,#2;{\\panelwidth=#1 \\panelheight=#2} \\def\\setbb #1,#2;#3,#4;#5,#6;{% \\tbxmin=#1 \\tbymin=#2 % \\bxmin=#3 \\bymin=#4 % \\bxmax=#5 \\bymax=#6} % \\def\\barepanel #1{% \\ifnum\\panelheight=0 \\tpdif=\\bymax \\advance\\tpdif by -\\bymin \\multiply \\tpdif by \\panelwidth \\tpanelheight=\\tpdif \\tpdif=\\bxmax \\advance\\tpdif by -\\bxmin \\divide \\tpanelheight by \\tpdif \\else \\tpanelheight=\\panelheight \\fi \\epsfig{file=#1,% bbllx=\\bxmin bp,bblly=\\bymin bp,bburx=\\bxmax bp,bbury=\\bymax bp,clip=,% width=\\panelwidth mm,height=\\tpanelheight mm}} \\def\\labelypanel #1{% \\ifnum\\panelheight=0 \\tpdif=\\bymax \\advance\\tpdif by -\\bymin \\multiply \\tpdif by \\panelwidth \\tpanelheight=\\tpdif \\tpdif=\\bxmax \\advance\\tpdif by -\\bxmin \\divide \\tpanelheight by \\tpdif \\else \\tpanelheight=\\panelheight \\fi \\tpdif=\\bxmax \\advance\\tpdif by -\\tbxmin \\tpanelwidth=\\panelwidth \\multiply \\tpanelwidth by \\tpdif \\tpdif=\\bxmax \\advance\\tpdif by -\\bxmin \\divide \\tpanelwidth by \\tpdif \\epsfig{file=#1,% bbllx=\\tbxmin bp,bblly=\\bymin bp,bburx=\\bxmax bp,bbury=\\bymax bp,% clip=,width=\\tpanelwidth mm,height=\\tpanelheight mm}} \\def\\labelxpanel #1{% \\ifnum\\panelheight=0 \\tpdif=\\bymax \\advance\\tpdif by -\\bymin \\multiply \\tpdif by \\panelwidth \\tpanelheight=\\tpdif \\tpdif=\\bxmax \\advance\\tpdif by -\\bxmin \\divide \\tpanelheight by \\tpdif \\else \\tpanelheight=\\panelheight \\fi \\tpdif=\\bymax \\advance\\tpdif by -\\tbymin \\multiply \\tpanelheight by \\tpdif \\tpdif=\\bymax \\advance\\tpdif by -\\bymin \\divide \\tpanelheight by \\tpdif \\epsfig{file=#1,% bbllx=\\bxmin bp,bblly=\\tbymin bp,bburx=\\bxmax bp,bbury=\\bymax bp,% clip=,width=\\panelwidth mm,height=\\tpanelheight mm}} \\def\\labelxypanel #1{% \\ifnum\\panelheight=0 \\tpdif=\\bymax \\advance\\tpdif by -\\bymin \\multiply \\tpdif by \\panelwidth \\tpanelheight=\\tpdif \\tpdif=\\bxmax \\advance\\tpdif by -\\bxmin \\divide \\tpanelheight by \\tpdif \\else \\tpanelheight=\\panelheight \\fi \\tpdif=\\bxmax \\advance\\tpdif by -\\tbxmin \\tpanelwidth=\\panelwidth \\multiply \\tpanelwidth by \\tpdif \\tpdif=\\bxmax \\advance\\tpdif by -\\bxmin \\divide \\tpanelwidth by \\tpdif \\tpdif=\\bymax \\advance\\tpdif by -\\tbymin \\multiply \\tpanelheight by \\tpdif \\tpdif=\\bymax \\advance\\tpdif by -\\bymin \\divide \\tpanelheight by \\tpdif \\epsfig{file=#1,% bbllx=\\tbxmin bp,bblly=\\tbymin bp,bburx=\\bxmax bp,bbury=\\bymax bp,% clip=,width=\\tpanelwidth mm,height=\\tpanelheight mm}} \\newcommand{\\panellabel}[3]{\\makebox[0pt][r]{\\raisebox{#2} {#3} \\hspace{#1}}} \\def\\topfraction{0.99} % \\def\\dbltopfraction{0.99} \\def\\textfraction{0.01} \\def\\floatpagefraction{0.99} \\def\\CC{\\par \\vspace*{-2ex} \\footnotesize \\baselineskip=8pt \\begin{verbatim}} \\def\\EE{\\par \\vspace*{-2ex} \\normalsize} \\long\\def\\startignore #1\\stopignore{} % \\newenvironment{task}{% \\begin{list}{$\\bullet$}{\\itemsep=1ex \\parindent=3ex \\parskip=1ex \\topsep=0ex}} {\\end{list}} \\def\\koprr{\\list{}{ \\itemindent=0pt \\listparindent=0pt \\itemsep=0pt \\leftmargin=5ex}\\item[]} % \\let\\endkoprr=\\endlist \\def\\kolom{\\list{}{\\rightmargin=5cm % \\topsep=0ex \\itemsep=0cm \\parsep=0ex \\parskip=0ex \\partopsep=0ex \\listparindent=3ex \\itemindent=0pt \\leftmargin=0pt} \\item[] \\vspace*{-0ex}} % \\let\\endkolom=\\endlist \\renewcommand{\\labelitemi}{{\\bf --}} % \\def\\setlistparams{ \\topsep=0.7ex % \\itemsep=0.7ex % \\leftmargini=3ex} % \\setlistparams % \\newcounter{alistindex} \\newcounter{alistvalue} % \\newenvironment{alist}{% \\begin{list}{\\alph{alistindex})}{\\usecounter{alistindex}} \\itemsep=1.5ex \\parskip=1ex}{\\end{list}} \\def\\spacing #1{\\small \\renewcommand{\\baselinestretch}{#1} \\normalsize \\setlistparams} % \\newenvironment{itemrr}{\\vspace*{-\\parskip} \\leftmargini=3ex \\topsep=0.6ex \\begin{itemize} \\itemsep=0.5ex \\parindent=0ex \\parsep=0ex}{ \\end{itemize} \\vspace*{-\\parskip}} \\newenvironment{conditem}{ \\leftmargini=6ex \\topsep=0.2ex \\begin{itemize} \\itemsep=0ex \\parindent=3ex \\parsep=0ex}{ \\end{itemize}} \\newenvironment{enumerr}{ \\leftmargini=3ex \\topsep=1.5ex \\begin{enumerate} \\itemsep=0.7ex \\parindent=3ex \\parsep=0ex}{ \\end{enumerate}} \\newenvironment{condenum}{ \\leftmargini=3ex \\topsep=0.5ex \\begin{enumerate} \\itemsep=0ex \\parindent=3ex \\parsep=0ex}{ \\end{enumerate}} \\newcounter{romenumnr} \\newenvironment{romanenum}{ \\leftmargini=5ex \\topsep=1.5ex \\begin{list}{(\\roman{romenumnr}) --}{\\usecounter{romenumnr} \\itemsep=1ex \\listparindent=3ex \\parsep=0ex}}{ \\end{list}} \\newcommand{\\posterpage}[2]{{\\Large \\sf \\newpage \\setlength{\\minipagewidth}{\\textwidth} \\addtolength{\\minipagewidth}{-2cm} \\fboxrule=0.3ex \\fboxsep=0ex \\def\\shadowparams{\\topsep=-2.5ex \\itemsep=-1.5ex \\leftmargini=5ex} \\renewcommand{\\labelitemi}{$\\bullet$} \\parskip=3ex \\shadowparams \\savebox{\\boxcontent}[\\textwidth]{ \\begin{minipage}[t]{\\minipagewidth} \\parskip=3ex \\vspace*{1ex} \\begin{center} \\LARGE \\bf #1 \\end{center} \\vspace{2ex} #2 \\vspace*{3ex} \\mbox{ } \\end{minipage}} \\hspace*{\\textwidth} \\hspace*{-0.8ex} \\rule{2ex}{\\dp\\boxcontent} \\par \\vspace*{-\\dp\\boxcontent} \\vspace*{-7.2ex} \\framebox[\\textwidth]{\\usebox{\\boxcontent}} \\par \\vspace*{-3.0ex} \\hspace*{1.7ex} \\rule{\\textwidth}{2ex} \\setlistparams \\renewcommand{\\labelitemi}{{\\bf --}}}} \\newcommand{\\vwgraph}[3]{\\newpage \\begin{center} \\shadowframe{#1}{ \\begin{center} \\vspace*{2ex} {\\Large \\bf #2}\\\\[4ex] \\end{center} {\\Large \\sf #3}} % \\end{center}} \\newenvironment{itemvg}{ \\leftmargini=3ex \\topsep=0.5ex \\begin{itemize} \\vspace{-1ex} \\itemsep=0.5ex \\parindent=3ex \\parsep=0ex}{ \\end{itemize}} \\long\\def\\lectvw#1#2{\\newpage \\vspace*{-15mm} \\begin{center} \\begin{center} {\\LARGE \\sc \\underline{#1}}\\\\[4ex] \\end{center} {\\Large \\sf #2} % \\end{center}} \\newlength{\\minipagewidth} \\newcommand{\\shadowframe}[2]{ \\setlength{\\minipagewidth}{#1} \\addtolength{\\minipagewidth}{-1cm} \\fboxrule=0.5ex \\fboxsep=0ex \\def\\shadowparams{\\topsep=-2.5ex \\itemsep=-1.5ex \\leftmargini=5ex} \\renewcommand{\\labelitemi}{$\\bullet$} \\parskip=3ex \\shadowparams \\savebox{\\boxcontent}[#1]{ \\begin{minipage}[t]{\\minipagewidth} \\vspace*{1.5ex} #2 \\vspace*{0.5ex} \\mbox{ } \\end{minipage}} \\hspace*{#1} \\hspace*{1.0ex} \\rule{2ex}{\\dp\\boxcontent} \\par \\vspace*{-\\dp\\boxcontent} \\vspace*{-7.9ex} \\framebox[#1]{\\usebox{\\boxcontent}} \\par \\vspace*{-3.0ex} \\hspace*{4.5ex} \\rule{#1}{2ex} \\setlistparams \\renewcommand{\\labelitemi}{{\\bf --}}} \\newsavebox{\\boxcontent} \\newcommand{\\ovalhead}[1]{ \\unitlength=1cm \\sbox{\\boxcontent}{\\mbox{~~{#1}~~}} \\begin{center} \\ifdim\\wd\\boxcontent>6ex \\ifdim\\wd\\boxcontent<8cm \\begin{picture}(8,3) \\thicklines \\put(4.0,0.8){\\oval(8,1.6)} \\put(0.0,0.7){\\parbox{8cm}{ \\begin{center} \\usebox{\\boxcontent} \\end{center}}} \\end{picture} \\else \\ifdim\\wd\\boxcontent<12cm \\begin{picture}(12,3) \\thicklines \\put(6.0,0.8){\\oval(12,1.6)} \\put(0.0,0.7){\\parbox{12cm}{ \\begin{center} \\usebox{\\boxcontent} \\end{center}}} \\end{picture} \\else \\begin{picture}(16,3) \\thicklines \\put(8.0,0.8){\\oval(16,1.6)} \\put(0.0,0.7){\\parbox{16cm}{ \\begin{center} \\usebox{\\boxcontent} \\end{center}}} \\end{picture} \\fi \\fi \\fi \\end{center}} \\newcommand{\\colloqframe}[2]{ \\setlength{\\minipagewidth}{#1} \\addtolength{\\minipagewidth}{-1cm} \\fboxrule=0.5ex \\fboxsep=0ex \\parskip=3ex \\savebox{\\boxcontent}[#1]{ \\begin{minipage}[t]{\\minipagewidth} \\vspace*{1.5ex} #2 \\vspace*{0.5ex} \\mbox{ } \\end{minipage}} \\hspace*{#1} \\hspace*{-0.8ex} \\rule{2ex}{\\dp\\boxcontent} \\par \\vspace*{-\\dp\\boxcontent} \\vspace*{-7.9ex} \\framebox[#1]{\\usebox{\\boxcontent}} \\par \\vspace*{-3.0ex} \\hspace*{1.7ex} \\rule{#1}{2ex} \\renewcommand{\\labelitemi}{{\\bf --}}} \\newcommand{\\soloval}[1]{\\fboxrule=0.5mm \\unitlength=1cm \\hspace*{5mm} \\sol{20mm}\\\\[-20mm] \\begin{picture}(15,3) \\thicklines \\put(7.5,2){\\oval(15.60,3.60)} \\put(7.5,2){\\oval(15.55,3.55)} \\put(7.5,2){\\oval(15.50,3.50)} \\put(7.5,2){\\oval(15.45,3.45)} \\put(7.5,2){\\oval(15.40,3.40)} \\put(7.5,2){\\oval(15.35,3.35)} \\put(7.5,2){\\oval(15.00,3.00)} \\put(2.0,1.9){\\parbox{13cm}{#1}} \\end{picture}} \\setcounter{secnumdepth}{3} \\setcounter{tocdepth}{3} \\def\\chapterrr#1{\\chapter{#1} \\label{chap:#1} \\vskip-\\parskip} \\def{} {We aim to determine the fundamental parameters of a sample of B stars with apparent visual magnitudes below 8 in the field-of-view of the CoRoT space mission, from high-resolution spectroscopy. } {We developed an automatic procedure for the spectroscopic analysis of B-type stars with winds, based on an extensive grid of FASTWIND model atmospheres. We use the equivalent widths and/or the line profile shapes of continuum normalized hydrogen, helium and silicon line profiles to determine the fundamental properties of these stars in an automated way. } {After thorough tests, both on synthetic datasets and on very high-quality, high-resolution spectra of B stars for which we already had accurate values of their physical properties from alternative analyses, we applied our method to 66 B-type stars contained in the ground-based archive of the CoRoT space mission. We discuss the statistical properties of the sample and compare them with those predicted by evolutionary models of B stars.} {Our spectroscopic results provide a valuable starting point for any future seismic modelling of the stars, should they be observed by CoRoT.} ", "introduction": "The detailed spectroscopic analysis of B-type stars has for a long time been restricted to a limited number of targets. Reasons for this are the a priori need for a realistic atmosphere model, the lack of large samples with high-quality spectra, and the long-winded process of line profile fitting, as the multitude of photospheric and wind parameters requires a large parameter space to be explored. The advent of high-resolution, high signal-to-noise spectroscopy in the nineties led to a renewed interest of the scientific community in spectroscopic research, and in particular in the relatively poorly understood massive stars. The establishment of continuously better instrumentation and the improvement in quality of the obtained spectroscopic data triggered a series of studies, which led to a rapid increase in our knowledge of massive stars. In this respect, it is not surprising to note that this is exactly the period where several groups started to upgrade their atmosphere prediction code for such stars, see, e.g., CMFGEN - \\citet{Hillier1998}, PHOENIX - \\citet{Hauschildt1999}, WM-Basic - \\citet{Pauldrach2001}, POWR - \\citet{Grafener2002} and FASTWIND - \\citet{Santolaya1997}, \\citet{Puls2005}. Initially, major attention was devoted to the establishment of a realistic atmosphere model (improvement of atomic data, inclusion of line blanketing and clumping), rather than analyzing large samples of stars. Simultaneously with improvements in the atmosphere predictions, also the number of available high-quality data increased rapidly, mainly thanks to the advent of multi-object spectroscopy. At the time of writing, the largest survey is the VLT-FLAMES Survey of Massive Stars \\citep{Evans2005}, containing over 600 Galactic, SMC and LMC B-type spectra\\footnote{additionally, roughly 90 O-stars have been observed.} (in 7 different clusters), gathered over more than 100 hours of VLT time. The survey not only allowed to derive the stellar parameters and rotational velocities for hundreds of stars \\citep{Dufton2006, Hunter2008}, but also to study the evolution of surface N abundances and the effective temperature scales in the Galaxy and Magellanic Clouds \\citep{Trundle2007, Hunter2007, Hunter2008a}. In preparation of the CoRoT space mission, and almost contemporary with the FLAMES setup, another large database was constructed: GAUDI (Ground-based Asteroseismology Uniform Database Interface, \\citealt{Solano2005}). It gathers ground-based observations of more than 1500 objects, including high-resolution spectra of about 250 massive B-type stars, with the goal to determine their fundamental parameters as input for seismic modeling (see Section\\,\\ref{GAUDI}). The availability of such large samples of B-type stars brings within reach different types of studies, e.g., they may lead to a significant improvement in the fundamental parameter calibration for this temperature range and to a confrontation with and evaluation of stellar evolution models (e.g., \\citealt{Hunter2008}). The drawback of this huge flood of data, however, is, as mentioned before, the large parameter space to be explored, which can be quite time-consuming, if no adequate method is available. It requires a method which is able to derive the complete set of parameters of stars with a wide variety of physical properties in an objective way. To deal with the large GAUDI dataset, we investigated the possibility of \\textit{automated} spectral line fitting and we opted for a grid-based fitting method: AnalyseBstar. In Section\\,\\ref{method}, we justify our choice for a grid-based method, present its design and discuss several tests which were applied to check the performance of the routine. A more detailed description of our methodology can be found in the (online) appendix. Section\\,\\ref{GAUDI} illustrates the first application of AnalyseBstar to the sample of CoRoT candidate targets in the GAUDI database and Section\\,\\ref{interpretation} deals with the physical interpretation and some statistical properties of the resulting parameters. Section\\,\\ref{summary} summarizes the main results obtained in this paper. ", "conclusions": "" }, "0910/0910.2256_arXiv.txt": { "abstract": "{ The lowest-mass stars, brown dwarfs and extrasolar planets present challenges and opportunities for understanding dynamics and cloud formation processes in low-temperature atmospheres. For brown dwarfs, the formation, variation and rapid depletion of photospheric clouds in L- and T-type dwarfs, and spectroscopic evidence for non-equilibrium chemistry associated with vertical mixing, all point to a fundamental role for dynamics in vertical abundance distributions and cloud/grain formation cycles. For exoplanets, azimuthal heat variations and the detection of stratospheric and exospheric layers indicate multi-layered, asymmetric atmospheres that may also be time-variable (particularly for systems with highly elliptical orbits). Dust and clouds may also play an important role in the thermal energy balance of exoplanets through albedo effects. For all of these cases, 3D atmosphere models are becoming an increasingly essential tool for understanding spectral and temporal properties. In this review, I summarize the observational evidence for clouds and dynamics in cool dwarf and hot exoplanetary atmospheres, outstanding problems associated with these processes, and areas where effective synergy can be achieved. ", "introduction": "Fifteen years ago, the first examples of brown dwarfs and extrasolar planets were discovered, sources which continue to challenge our understanding of low temperature atmospheres. Several hundred very low-mass stars and brown dwarfs are now known (the latter distinguished by their lack of core hydrogen fusion), encompassing two newly defined spectral classes, the L dwarfs and the T dwarfs (see review by \\citealt{2005ARA&A..43..195K}). Roughly 300 extrasolar planets have been found through Doppler shift, transit, microlensing and direct detection techniques. Collectively, these low-temperature sources span a broad range of mass (1~{\\mjup} $\\lesssim$ M $\\lesssim$ 100~{\\mjup}), photospheric temperature (100~K $\\lesssim$ $T$ $\\lesssim$ 2500~K), surface gravity (3 $\\lesssim$ {\\logg} $\\lesssim$ 5.5~cm~s$^{-2}$), age (few Myr to few Gyr), elemental composition, rotation period (hours to days), magnetic activity, degree of external heating and interior structure. Yet all are related by their cool, molecule-rich, dynamic atmospheres. While direct studies of cool dwarf atmospheres have been feasible since their discovery, it is only recently that techniques to probe exoplanet atmospheres have been realized. The best-constrained exoplanets are those which closely orbit and transit their host stars---so-called ``Hot Jupiters''---allowing reflectance and thermal spectrophotometry during secondary transit (e.g., \\citealt{2005ApJ...626..523C, 2005Natur.434..740D}), and transmission spectrophotometry during primary transit (e.g., \\citealt{2002ApJ...568..377C}). Phase curves for non-transiting systems have also been measured (e.g., \\citealt{2006Sci...314..623H}), and the recent direct detection of planets around the young stars Fomalhaut, $\\beta$ Pictoris and HR~8799 \\citep{2008Sci...322.1345K,2008Sci...322.1348M, 2009A&A...493L..21L} through high contrast imaging techniques have opened the door to direct investigations of exoplanet atmospheres. Two themes are prominent in the interpretation of observational data for low-temperature stellar, brown dwarf and exoplanetary atmospheres: clouds and dynamics. Condensate clouds are an important source of opacity and a component in the chemical network; they also modulate the albedos, energy budgets and atmospheric structure of exoplanets under intense irradiation. Dynamics are responsible for the redistribution of heat and modification of chemical abundances, and are likely fundamental to cloud formation and evolution processes. In this review, I summarize our current observational evidence for clouds and dynamics in cool dwarfs and hot exoplanets, and note outstanding issues in their role in emergent spectral energy distributions, long-term thermal evolution, albedo, and temporal and azimuthal variability. I conclude with a discussion of opportunities for synergy in future generations of cool dwarf and exoplanet models, with a view toward 3D simulations. ", "conclusions": "" }, "0910/0910.1533_arXiv.txt": { "abstract": "The stellar populations of galaxies contain a wealth of detailed information. From the youngest, most massive stars, to almost invisible remnants, the history of star formation is encoded in the stars that make up a galaxy. Extracting some, or all, of this information has long been a goal of stellar population studies. This was achieved in the last couple of decades and it is now a routine task, which forms a crucial ingredient in much of observational galaxy evolution, from our Galaxy out to the most distant systems found. In many of these domains we are now limited not by sample size, but by systematic uncertainties and this will increasingly be the case in the future. The aim of this review is to outline the challenges faced by stellar population studies in the coming decade within the context of upcoming observational facilities. I will highlight the need to better understand the near-IR spectral range and outline the difficulties presented by less well understood phases of stellar evolution such as thermally pulsing AGB stars, horizontal branch stars and the very first stars. The influence of rotation and binarity on stellar population modelling is also briefly discussed. ", "introduction": "\\label{sec:introduction} The luminous output of normal galaxies is ultimately generated by stars in various stages of stellar evolution --- in isolation a trite observation, but this simple fact gives us the opportunity to extract a vast amount of information from observations of galaxies through the modelling of their stellar populations. While the study of stellar populations in galaxies started with Baade's (1944)\\nocite{1944ApJ...100..137B} identification of two populations of stars in M32 and NGC 205, a rigourous study of the topic only commenced in the late 60's and early 70's \\citep{Tinsley1968,1972A&A....20..361F,1973ApJ...179..427S} with Tinsley's \\textit{Fundamentals of Cosmic Physics} article \\citep{1980FCPh....5..287T} particularly influential. Since that time, the number of articles discussing stellar populations has risen rapidly so that today about 12\\% of all articles in the major journals mention stellar populations in their abstracts (Figure 1). The majority of this growth has been made possible through the development of simple models for the evolution of stellar populations that have found widespread use in a wide range of astronomical studies, from stellar clusters in the Milky Way to the most distant galaxies in the Universe. I will repeatedly refer to stellar population models below, and use this term loosely to refer to any model that predict the observational properties of any ensemble of stars. These start from models of stellar evolution (e.g. Bertelli et al 1994; see contribution by Cassisi these proceedings). By applying empirical colour corrections (e.g. Lejeune et al 1997) they can be placed on an observational Hertzsprung-Russell diagram. For a given star formation history and initial mass function (IMF), this can be sampled, either to produce a Monte Carlo realisation of an observed colour-magnitude diagrams (see Tolstoy et al 2009), or by convolution to create integrated properties of a stellar population (e.g. Fioc \\& Rocca-Volmerange 1997; Leitherer et al 1999; Vazdekis 1999; Bruzual \\& Charlot 2003; Maraston 2005; Kotulla et al 2009). Finally, these models are compared to observations using a number of techniques, in itself an interesting topic, but for this review the details of this step are not crucial. In this contribution I will begin by briefly reviewing some of the achievements in extra-galactic astronomy made possible by our understanding of stellar populations, before I turn to the challenges for stellar populations studies in the coming decade. Given the format this will not be an exhaustive review, see the other contributions in these proceedings for more in-depth discussion of many of the topics. ", "conclusions": "\\label{sec:conclusions} Rather than a traditional conclusion, I would here like to end by summarising the challenges above into a set of rest wavelength ranges, highlighting the issues and some suggestions as to what might be useful studies to do --- note that most of these suggestions are merely reinforcing already existing studies! \\textbf{Far-UV} --- \\textit{Massive stars --- what are realistic rotation and binary parameters}. \\textbf{To do:} Observations of lensed Lyman-break galaxies, COS observations of nearby stars and star-forming regions, more in-depth studies of stellar mass-loss \\textbf{Near-UV} --- \\textit{Binary evolution, mass loss on the RGB}. \\textbf{To do:} COS and GALEX spectroscopy of UV-upturn galaxies, inclusion of horizontal branch uncertainties in models. Combination of optical, near-UV and X-ray data in analysis. \\textbf{Optical} --- \\textit{High-precision predictions. Non-solar abundance variations and impact on spectra, emission lines}. \\textbf{To do:} Careful and extensive testing of theoretical spectra, comparison to detailed spectroscopic analysis of nearby systems: what abundances can be reliably extracted from medium resolution spectra? \\textbf{Near-IR} --- \\textit{Evolution of and spectra of luminous, cold stars. TP-AGB stars, RGs and RSGs, spectral features in the near-IR}. \\textbf{To do:} Further observations of resolved stellar populations in the near-IR with and without adaptive optics with careful comparison to results including optical data. Calibration spectral features in the near-IR. Obtain ACS optical data to back up future studies in near-IR with ELTs. In depth analysis of uncertainties in population synthesis models and careful comparisons to data in the nearby and distant Universe. \\begin{multicols}{2}" }, "0910/0910.5917_arXiv.txt": { "abstract": "The presence of convective motions in the atmospheres of metal-poor halo stars leads to systematic asymmetries of the emergent spectral line profiles. Since such line asymmetries are very small, they can be safely ignored for standard spectroscopic abundance analysis. However, when it comes to the determination of the $^6$Li/$^7$Li isotopic ratio, $q$(Li)=$n$($^6$Li)/n($^7$Li), the intrinsic asymmetry of the $^7$Li line must be taken into account, because its signature is essentially indistinguishable from the presence of a weak $^6$Li blend in the red wing of the $^7$Li line. In this contribution we quantity the error of the inferred $^6$Li/$^7$Li isotopic ratio that arises if the convective line asymmetry is ignored in the fitting of the $\\lambda\\,6707$~\\AA\\ lithium blend. Our conclusion is that $^6$Li/$^7$Li ratios derived by \\cite[Asplund \\etal\\ (2006)]{A2006}, using symmetric line profiles, must be reduced by typically $\\Delta q$(Li) $\\approx 0.015$. This diminishes the number of certain $^6$Li detections from 9 to 4 stars or less, casting some doubt on the existence of a $^6$Li plateau. ", "introduction": "The spectroscopic signature of the presence of $^6$Li in the atmospheres of metal-poor halo stars is a subtle extra depression in the red wing of the $^7$Li doublet, which can only be detected in spectra of the highest quality. Based on high-resolution, high signal-to-noise VLT/UVES spectra of 24 bright metal-poor stars, \\cite[Asplund \\etal\\ (2006)]{A2006} report the detection of $^6$Li in nine of these objects. The average $^6$Li/$^7$Li isotopic ratio in the nine stars in which $^6$Li has been detected is $q$(Li) $\\approx 0.04$ and is very similar in each of these stars, defining a $^6$Li plateau at approximately $\\log n(^6$Li$) = 0.85$ (on the scale $\\log n($H$) = 12$). A convincing theoretical explanation of this new $^6$Li plateau turned out to be problematic. Even when the depletion of the $^6$Li isotope during the pre-main-sequence phase would be ignored, the high abundances of $^6$Li at the lowest metallicities cannot be explained by current models of galactic cosmic-ray production (for a concise review see e.g.\\ \\cite[Christlieb 2008]{C2008}, and references therein). A possible solution of the so-called `second Lithium problem' was suggested by \\cite[Cayrel \\etal\\ (2007)]{C2007}, who point out that the intrinsic line asymmetry caused by convection in the photospheres of cool stars is almost indistinguishable from the asymmetry produced by a weak $^6$Li blend on a presumed symmetric $^7$Li profile. As a consequence, the derived $^6$Li abundance should be significantly reduced when the intrinsic line asymmetry in properly taken into account. Using 3D non-LTE line formation calculations based on 3D hydrodynamical model atmospheres computed with the CO$^5$BOLD code (\\cite[Freytag \\etal\\ 2002]{F2002}, \\cite[Wedemeyer \\etal\\ 2004]{W2004}, see also {\\tt http://www.astro.uu.se/$\\sim$bf/co5bold\\_main.html}), we quantify the theoretical effect of the convection-induced line asymmetry on the resulting $^6$Li abundance as a function of effective temperature, gravity, and metallicity, for a parameter range that covers the stars of the \\cite[Asplund \\etal\\ (2006)]{A2006} sample. ", "conclusions": "The present study indicates that only $2$ or at most $4$ out of the $24$ stars of the \\cite{A2006} sample remain significant $^6$Li detections when subjected to a 3D~non-LTE analysis, suggesting that the presence of $^6$Li in the atmospheres of galactic halo stars is rather the exception than the rule. This would imply that it is no longer necessary to look for a global mechanism accounting for a $^6$Li enrichment of the galactic halo, but that it is sufficient to explain only a few exceptional cases, which is probably much easier.\\\\[-5mm]" }, "0910/0910.2706_arXiv.txt": { "abstract": "Many astrophysical sources, especially compact accreting sources, show strong, random brightness fluctuations with broad power spectra in addition to periodic or quasi-periodic oscillations (QPOs) that have narrower spectra. The random nature of the dominant source of variance greatly complicates the process of searching for possible weak periodic signals. We have addressed this problem using the tools of Bayesian statistics; in particular using Markov chain Monte Carlo techniques to approximate the posterior distribution of model parameters, and posterior predictive model checking to assess model fits and search for periodogram outliers that may represent periodic signals. The methods developed are applied to two example datasets, both long \\xmm\\ observations of highly variable Seyfert 1 galaxies: RE J$1034+396$ and Mrk $766$. In both cases a bend (or break) in the power spectrum is evident. In the case of RE J$1034+396$ the previously reported QPO is found but with somewhat weaker statistical significance than reported in previous analyses. The difference is due partly to the improved continuum modelling, better treatment of nuisance parameters, and partly to different data selection methods. ", "introduction": "\\label{sect:intro} \\defcitealias{Vaughan05}{V05} A perennial problem in observational astrophysics is detecting periodic or almost-periodic signals in noisy time series. The standard analysis tool is the periodogram \\citep[see e.g. ][]{Jenkins69, Priestley81, Press92, Bloomfield00, Chatfield03}, and the problem of period detection amounts to assessing whether or not some particular peak in the periodogram is due to a periodic component or a random fluctuation in the noise spectrum \\citep[see][]{Fisher29, Priestley81, Leahy83, vanderklis89, Percival93, Bloomfield00}. If the time series is the sum of a random (stochastic) component and a periodic one we may write $y(t) = y_R(t) + y_P(t)$ and, due to the independence of $y_R(t)$ and $y_P(t)$, the power spectrum of $y(t)$ is the sum of the two power spectra of the random and stochastic processes: $S_Y(f) = S_R(f) + S_P(f)$. This is a \\emph{mixed} spectrum \\citep[][section 4.4]{Percival93} formed from the sum of $S_P(f)$, which comprises only narrow features, and $S_R(f)$, which is a continuous, broad spectral function. Likewise, we may consider an evenly sampled, finite time series $y(t_i)$ ($i=1,2,\\ldots,N$) as the sum of two finite time series: one is a realisation of the periodic process, the other a random realisation of the stochastic process. We may compute the periodogram (which is an estimator of the true power spectrum) from the squared modulus of the Discrete Fourier Transform (DFT) of the time series, and, as with the power spectra, the periodograms of the two processes add linearly: $I(f_j) = I_R(f_j) + I_P(f_j)$. The periodogram of the periodic time series will contain only narrow ``lines'' with all the power concentrated in only a few frequencies, whereas the periodogram of the stochastic time series will show power spread over many frequencies. Unfortunately the periodogram of stochastic processes fluctuates wildly around the true power spectrum, making it difficult to distinguish random fluctuations in the noise spectrum from truly spectral periodic components. See \\cite{vanderklis89} for a thorough review of these issues in the context of X-ray astronomy. Particular attention has been given to the special case that the spectrum of the stochastic process is flat (a \\emph{white noise} spectrum $S(f) = const$), which is the case when the time series data $y_R(t_i)$ are independently and identically distributed (IID) random variables. Reasonably well-established statistical procedures have been developed to help identify spurious spectral peaks and reduce the chance of false detections \\citep[e.g.][]{Fisher29, Priestley81, Leahy83, vanderklis89, Percival93}. In contrast there is no comparably well-established procedure in the general case that the spectrum of the stochastic process is not flat. In a previous paper, \\cite{Vaughan05} (henceforth \\citetalias{Vaughan05}), we proposed what is essentially a generalisation of Fisher's method to the case where the noise spectrum is a power law: $S_R(f) = \\beta f^{-\\alpha}$ (where $\\alpha$ and $\\beta$ are the power law index and normalisation parameters). Processes with power spectra that show a power law dependence on frequency with $\\alpha > 0$ (i.e. increasing power to lower frequencies) are called \\emph{red noise} and are extremely common in astronomy and elsewhere \\citep[see][]{Press78}. In this paper we expand upon the ideas in \\citetalias{Vaughan05} and, in particular, address the problem from a Bayesian perspective that allows further generalisation of the spectral model of the noise. The rest of this paper is organised as follows. In section~\\ref{sect:bayes} we introduce some of the basic concepts of the Bayesian approach to statistical inference; readers familiar with this topic may prefer to skip this section. Section~\\ref{sect:stat} gives a brief overview of classical significance testing using $p$-values (tail area probabilities) and test statistics, and section~\\ref{sect:ppp} discusses the posterior predictive $p$-value, a Bayesian counterpart to the classical $p$-value. Section \\ref{sect:pc} reviews the conventional (classical) approaches to testing for periodogram peaks. Section \\ref{sect:ml} outlines the theory of maximum likelihood estimation from periodogram data, which is developed into the basis of a fully Bayesian analysis in sections~\\ref{sect:ba} and \\ref{sect:ppper}. The Bayesian method is then applied to two real observations if AGN in section \\ref{sect:data}. Section \\ref{sect:disco} discusses the limitations of the method, and alternative approaches to practical data analysis. A few conclusions are given in section \\ref{sect:conc}, and two appendices describe details of the simulations algorithms used in the analysis. ", "conclusions": "\\label{sect:conc} We have presented Bayesian methods for the modelling of periodogram data that can be used for both parameter estimation and model checking, and may be used to test for narrow spectral features embedded in noisy data. The model assessment is performed using simulations of posterior predictive data to calibrate (sensibly chosen) test statistics. This does however leave some arbitrariness in the method, particularly in the choice of test statistic\\footnote{In situations where two competing models can be modelled explicitly the LRT provides a natural choice of statistic.} (and in some situations the choice of what constitutes a simulation of the data). Such issues were always present, if usually ignored, in the standard frequentist tests. The posterior predictive approach has the significant advantage of properly treating nuisance parameters, and provides a clear framework for checking the different aspects of the reasonableness of a model fit. The issue of choosing a test statistic does not arise in more ``purist'' Bayesian methods such as Bayes factors, which concentrate on the posterior distributions and marginal likelihoods, but such methods of model selection carry their own burden in terms of the computational complexity and the difficulty of selecting (and the sensitivity of inferences to) priors on the model parameters. The method presented in this paper, making use of the posterior predictive checking, is an improvement over the currently popular methods that use classical $p$-value; but Bayesian model selection is an area of active research and it is not unreasonable to expect that new, powerful and practical computational tools will be developed or adapted to help bridge the gap between the pragmatic and the purist Bayesian approaches. The routines used to perform the analysis of the real data presented in section \\ref{sect:data} will be made available as an {\\tt R}\\footnote{{\\tt R} is a powerful, open-source computing environment for data analysis and statistics that may be downloaded for free from {\\tt http://www.r-project.org/} \\citep{r, r2}.} script from the author upon request." }, "0910/0910.0159_arXiv.txt": { "abstract": "We investigate the case of [{\\sc Cii}] 158$\\mu$m observations for SPICA/SAFARI using a three-dimensional magnetohydrodynamical (MHD) simulation of the diffuse interstellar medium (ISM) and the Meudon PDR code. The MHD simulation consists of two converging flows of warm gas ($10^4$ K) within a cubic box 50 pc in length. The interplay of thermal instability, magnetic field and self-gravity leads to the formation of cold, dense clumps within a warm, turbulent interclump medium. We sample several clumps along a line of sight through the simulated cube and use them as input density profiles in the Meudon PDR code. This allows us to derive intensity predictions for the [{\\sc Cii}] 158$\\mu$m line and provide time estimates for the mapping of a given sky area. ", "introduction": "The fine structure line of ionised carbon [{\\sc Cii}] at 157.7~$\\mu$m is one of the most prominent far infrared (FIR) line. It is widespread in the Milky Way as revealed by the COBE-FIRAS observations, and detected in the diffuse ISM \\citep{levrierf:I02}, photo-dissociation regions as well as local and distant galaxies \\citep[and references therein]{levrierf:M09}. [{\\sc Cii}] is one of the main cooling lines of the neutral ISM, in all environments where carbon is mostly ionised, that is where UV-photons with $h\\nu \\geq 11.3~\\mathrm{eV}$ are available. Unfortunately, the present information on the spatial structure of the [{\\sc Cii}] emission is very limited, because of the small telescope size of previous far infrared missions. While Herschel and SOFIA will improve the situation for bright regions, their sensitivity will be limited for the extended emission from either the diffuse interstellar medium, or the outskirts of molecular clouds. The structure of the [{\\sc Cii}] emission is expected to reveal the underlying structure of the matter, and provide information on the mechanisms responsible for the fragmentation. In this paper, we explore the perspectives offered by SPICA for mapping the interstellar medium. ", "conclusions": "This work demonstrates the ability of SPICA/SAFARI to map large areas of the sky in the [{\\sc Cii}] line in a much shorter time than what would be possible with Herschel, making it clear that the extraordinary sensitivity of the proposed mission is a definitive asset regarding this type of project. On another level, the work presented here is part of an ongoing ASTRONET project dubbed STAR FORMAT, which is a German-French collaboration whose two main objectives are : 1/ Producing databases to publicize simulation results from MHD and PDR codes; 2/ Developing MHD, PDR and radiative transfer codes in an interoperable fashion, to allow for a much improved treatment of the physics of the ISM, especially regarding the formation of molecular clouds and dense cores. With respect to this project, our work shows that a truly three-dimensional PDR code is a must, to adequately deal with the complex geometries of the structures formed in the MHD simulations. The development of parallel computing schemes is also quite inevitable, and it should be noted that the PDR code can already be run on a grid. It is hoped that the present paper can be used as ground work for these developments." }, "0910/0910.0968_arXiv.txt": { "abstract": "The geoeffective magnetic cloud (MC) of 20 November 2003, has been associated to the 18 November 2003, solar active events in previous studies. In some of these, it was estimated that the magnetic helicity carried by the MC had a positive sign, as well as its solar source, active region (AR) NOAA 10501. In this paper we show that the large-scale magnetic field of AR 10501 had a negative helicity sign. Since coronal mass ejections (CMEs) are one of the means by which the Sun ejects magnetic helicity excess into the interplanetary space, the signs of magnetic helicity in the AR and MC should agree. Therefore, this finding contradicts what is expected from magnetic helicity conservation. However, using for the first time correct helicity density maps to determine the spatial distribution of magnetic helicity injection, we show the existence of a localized flux of positive helicity in the southern part of AR 10501. We conclude that positive helicity was ejected from this portion of the AR leading to the observed positive helicity MC. ", "introduction": "\\label{S-Intro} Magnetic helicity globally quantifies the signed amount of twist, writhe, and shear of the magnetic field in a given volume (see the review by \\inlinecite{Demoulin07} and references therein). Magnetic helicity plays an important role in magnetohydrodynamics (MHD) because it is one of the few global quantities which are conserved, even in resistive MHD on time scales shorter than the global diffusion time scale (\\opencite{Berger84}). Coronal mass ejections (CMEs) are expulsions of mass and magnetic field from the Sun. \\inlinecite{Rust94} and \\inlinecite{Low96} pointed out that one of the most important roles of CMEs is to carry away magnetic helicity from the Sun. Otherwise, because helicity dissipates very slowly and helicities of opposite sign are globally injected through the photosphere in each solar hemisphere without change of sign during consecutive cycles (\\opencite{Berger00}), helicity would accumulate continuously. A fraction of CMEs can be observed {\\it in situ} as magnetic clouds (MCs). An MC is characterized by lower proton temperature and higher magnetic field strength than the surrounding solar wind. Typically, the magnetic field vector shows a smooth and significant rotation across the cloud (\\opencite{Burlaga81}; \\opencite{Klein82}) indicating a helical (flux rope) magnetic structure, which clearly has non-zero helicity. Therefore, a measurable prediction is that the interplanetary MC must carry the same amount of helicity that was ejected from the solar source region. In particular, the signs of the magnetic helicity of the MC and of the solar region from which it originates should agree. As a first approach, the magnetic helicity sign of some structures in the solar atmosphere can be inferred from certain observed morphological features (see the review by \\inlinecite{Demoulin09} and references therein). These features include {\\it sunspot whorls} (handedness of the spiral patterns of chromospheric fibrils), {\\it filament barbs} (direction of the barbs relative to the orientation of the magnetic field), {\\it flare-ribbons} (forward/reverse `J-shape' observed in H$\\alpha$ and UV wavelengths), {\\it sigmoids} (normal or reverse `S-shaped' loops in soft X-ray observations), {\\it magnetic tongues} (angle formed by the magnetic inversion line relative to the AR axis in emerging ARs), {\\it coronal loops} (orientation of loops in EUV observations relative to the magnetic inversion line), {\\it vector magnetic field} (direction of sheared fields). These observational features can be used to qualitatively compare the magnetic helicity sign of an MC with that of its solar source once identified ({\\it e.g.} \\opencite{Subramanian01}; \\opencite{Schmieder05}; \\opencite{Ali07}). In a similar way, {\\it i.e.} directly from observations, the helicity sign of MCs can be estimated from the measured rotation of the vector magnetic field (see examples in Bothmer and Schwenn (1994, 1998)). Several studies have found that the magnetic helicity sign of the MC and that inferred from the morphological features of its source AR match ({\\it e.g.} \\opencite{RK94}; Bothmer and Schwenn, 1994, 1998; \\opencite{Marubashi97}; \\opencite{Ruzmaikin03}; \\opencite{Rust05}). Other studies have determined the helicity sign of the CME source region modeling the coronal magnetic field ({\\it e.g.} \\opencite{Yurchyshyn01}, \\citeyear{Yurchyshyn06}), in these cases also the source region and cloud helicity signs were in agreement. However, the comparison does not yield a complete agreement, since a few MCs seem to present a different helicity sign from that of their solar source \\cite{Leamon04}. Quantitative comparisons of the helicity involved in the solar ejection and that of the associated MC have recently been possible. In the interplanetary medium, these quantitative comparisons require either the modeling of {\\it in situ} magnetic field observations (see the reviews by \\inlinecite{Dasso05} and \\inlinecite{Nakwacki08}) or, in cases when the impact parameter is small and considering a local cylindrical geometry, the MC helicity can be directly quantified from the data \\cite{Dasso06}. In the solar atmosphere, at least two different methods, giving consistent results (\\opencite{Lim07}), allow the estimation of the amount of ejected helicity. One way is to compute the helicity variation before and after the ejection of the solar source region using a coronal field model (\\opencite{Green02}; \\opencite{Demoulin02}; \\opencite{Mandrini05}; \\opencite{Regnier05}; \\opencite{Luoni05}). This method requires the knowledge of the magnetic field in the entire volume. Another way is to measure the helicity injection, based on a time series of photospheric field observations. This was initiated by \\inlinecite{Chae01b} and \\inlinecite{Chae01}, and was subsequently applied to the study of CMEs by \\inlinecite{Nindos02} and \\inlinecite{Nindos03}. \\inlinecite{Pariat05} showed that, although the methods used in previous works could correctly estimate the total injected flux of helicity, they incorrectly determined the localized injected flux and proposed an alternative to properly map the helicity flux injection. We shall use this corrected method to compute the magnetic helicity injection in AR 10501. The large MC observed on 20 November 2003, gave place to the largest geomagnetic storm of solar cycle 23 \\cite{Gopalswamy05}. This MC was associated to the active solar events that occurred in AR 10501 on 18 November 2003, by several authors (\\opencite{Gopalswamy05}; \\opencite{Yurchyshyn05}; \\opencite{Mostl08}). The association discussed by \\inlinecite{Gopalswamy05} is mainly based on the timing between the filament ejections in AR 10501, the appearance of CMEs in the Large Angle and Spectroscopic Coronagraph (LASCO) white light images (its height-time plots), and the arrival of the MC to the Advanced Composition Explorer (ACE) spacecraft and its velocity. \\inlinecite{Yurchyshyn05} also estimated the AR helicity sign modeling the AR coronal magnetic field and the MC {\\it in situ} data; these latter authors showed that the helicity of the MC and AR were both positive. However, \\inlinecite{Mostl08} discussed that while the MC helicity sign seems well determined, the handedness (or helicity sign) of the very extended filament, lying along different portions of the inversion line within and in the surroundings of the AR, is ambiguous. In this paper, we revisit the evolution of the activity in AR 10501 along 18 November 2003, in Section~\\ref{S-obs}. Then, we discuss the characteristics of the positive-helicity-carrier MC observed by ACE on 20 November 2003, and its association with the CMEs originating from the AR (Section~\\ref{Ss-MC}). To verify this association, we analyze carefully the morphological features of the region; in particular, the different segments of filament material lying along the magnetic inversion line that are observed as forming a single and very extended filament. We find that all these observational features, except for one filament segment, indicate that the sign of the dominant helicity in the AR was negative. Global magnetic field extrapolations, before any AR activity, agree with this finding (Section~\\ref{S-Morph}). This is in strong contradiction with what is expected by models of MC formation constrained by the helicity conservation principle. In view of this result, we analyze the photospheric magnetic field evolution of the region and we perform an in-depth analysis of the local helicity flux injection along 18 November (Section~\\ref{S-LocHInj}). We find that a zone at the south of AR 10501 was the location of positive helicity flux injection during that day. Finally, we conclude that the ejected flux rope, observed later as an MC, should be located at this southern portion of the AR, and discuss the implications of our finding for CMEs/MCs triggering models (Section~\\ref{S-Concl}). ", "conclusions": "% \\label{S-Concl} We have determined the global and local magnetic helicity sign of AR 10501 on 18 November 2003, using different methods based on the analysis of a multi-wavelength data set. We have also discussed the association of this AR with the MC of 20 November 2003, observed by ACE, the largest geoeffective cloud during the solar cycle 23. AR 10501 is surrounded by a large apparent circular-shaped single filament, which is in fact formed by several distinct segments. The flares of 18 November 2003 were initiated by the continuous emergence of magnetic bipoles and by the eruption of some segments of the filament. The destabilization of one of the filament segments is the primary trigger of the third flare. As discussed in the `CSHKP' (\\opencite{Carmichael64}; \\opencite{Sturrock66}; \\opencite{Hirayama74}, and \\opencite{Kopp76}; also Forbes and Malherbe (1991)) standard solar-flare model, a filament eruption (in our particular case, the eruption of a filament segment) due to some magnetic instability occurs above the magnetic inversion line. As a result, filament material moves away from the solar surface, the pre-existing magnetic arcade stretches upward, and the condition arises for magnetic reconnection, leading to a CME ejection and a subsequent MC in the interplanetary medium. Based on the multi-wavelength data set, we have found the following results in relation to the magnetic helicity of the AR: \\smallskip \\noindent $-$ The very extended filament within and surrounding the AR presents evidence of dextral, as well as sinistral chirality, {\\it i.e.} negative and positive magnetic helicity. \\noindent $-$ The sunspot whorls show a left-hand twist, which corresponds to negative magnetic helicity. \\noindent $-$ The reverse `J-shaped' flare ribbons indicate negative helicity in the active region. However, this may be a false indicator because the ribbons follow the magnetic inversion line. \\noindent $-$ From the coronal magnetic field model, we have found that the best $\\alpha$ values to fit the large-scale TRACE loops are negative. These negative $\\alpha$ values mean that the magnetic helicity is globally negative. However, at the south of the AR, where N4/P4 lie, the value of $\\alpha$ needed to locally match the shape of loops in the `post-flare' arcade after the flare that peaked at 08:31 UT has a tendency to be positive towards the east and negative towards the west, in agreement with the location of sinistral and dextral filament segments. \\noindent $-$ The computation of the magnetic helicity injection, using G$_\\theta$ maps, indicates that the main spot in the AR is dominated by negative helicity. However, there is a strong local positive injection of helicity in the southern polarities (N4/P4) of the AR. \\smallskip From the above evidences, we conclude that AR 10501 has a global negative magnetic helicity. Despite this global negative helicity, the helicity density maps, {\\it i.e.} G$_\\theta$ maps, show a strong injection of positive magnetic helicity in the southern polarities. Our result provides a clear example of an AR in which the magnetic helicity sign is mixed, with simultaneous injection of both helicity signs. Previous works (\\opencite{Pevtsov97}; \\opencite{Green02}) have already reported examples in which the total helicity sign of an AR changed as it evolved because of parasitic magnetic polarity emergence having an opposite helicity sign to the main one. This is the first time that helicity flux density maps bring new information about the local helicity injection into an active region whose sign is opposite to the global helicity of the AR. Similar local injection of helicity of an opposite sign may explain the few cases of discrepancy found by \\inlinecite{Leamon04} between the helicity sign carried by MCs and the global helicity sign of their identified solar source region. We speculate that due to a global instability in the AR, a flux rope with positive helicity erupted. This flux rope would contain the filament segment with positive helicity. As discussed above, the erupting filament has sections with opposite chiralities (sinistral in F2a and dextral in F2b). Filaments with the same chirality can merge, while those with different chirality cannot merge (\\opencite{Martin98}; \\opencite{Rust01}; \\opencite{Schmieder04}; \\opencite{Aulanier06}). The segment F2 corresponds to two different flux tubes having the same axial magnetic field direction in agreement with the counterstreaming motions observed before the eruption. However, the two sections F2a and F2b have opposite chiralities. Possibly, during the third flare, these two segments interacted through magnetic reconnection and helicity carried by both section partially canceled. The positive helicity carried by F2a being larger, the CME and ICME induced by the eruption transported away a net positive helicity, as measured by ACE at 1 AU \\cite{Gopalswamy05}. Finally, our finding of a local injection of helicity with opposite signs is potentially important for eruption models. \\inlinecite{Kusano04} developed a model based on the reconnection of magnetic flux ropes with opposite helicity. Helicity annihilation potentially allows the release of a larger amount of free energy since the field can eventually relax to a state closer to potential. \\begin{acks} The authors thanks Dr. Pascal D\\'emoulin for fruitful discussions and suggestions. R.C. thanks the Le Centre Franco-Indien pour la Promotion de la Recherche Avanc\\'ee (CEFIPRA) for his postdoctoral grant. This work was done in the frame of the European network SOLAIRE. This work used the DAVE/DAVE4VM codes written and developed by the Naval Research Laboratory. E.P. wishes to thank Peter Schuck for providing the DAVE/DAVE4VM code and for useful discussions. The work of E.P. was supported, in part, by the NASA HTP and SR$\\&$T programs. C.H.M. thanks the Argentinean grants: UBACyT X127 and PICT 03-33370 (ANPCyT). C.H.M. is a member of the Carrera del Investigador Cient\\'{i}fico, CONICET. We acknowledge the use of TRACE data. MDI data are a courtesy of SOHO/MDI consortium. SOHO is a project of international cooperation between ESA and NASA. B.S and C.H.M have started this work in the frame of the ISSI workshop chaired by Dr. Consuelo Cid (2008-2010). We also thank the anonymous referee for helpful and constructive comments. \\end{acks}" }, "0910/0910.4529_arXiv.txt": { "abstract": "{Determination of the mass functions of open clusters of different ages allows us to infer the efficiency with which brown dwarfs are evaporated from clusters to populate the field.} {In this paper we present the results of a photometric survey to identify low mass and brown dwarf members of the old open cluster Praesepe (age 590$^{+150}_{-120}$\\,Myr, distance 190$^{+6.0}_{-5.8}$\\,pc) from which we estimate its mass function and compare this with that of other clusters.} {We performed an optical ($I_{\\rm c}$-band) and near-infrared ($J$ and $K_{\\rm s}$-band) photometric survey of Praesepe covering 3.1\\,deg$^2$. With 5$\\sigma$ detection limits of $I_{\\rm c}=23.4$ and $J=20.0$, our survey is predicted to be sensitive to objects with masses from 0.6 to 0.05\\,M$_\\odot$.} {We photometrically identify 123 cluster member candidates based on dust-free atmospheric models and 27 candidates based on dusty atmospheric models. The mass function rises from 0.6\\,M$_\\odot$ down to 0.1\\,M$_\\odot$ (a power law fit of the mass function gives $\\alpha$=1.8$\\pm$0.1; \\,$\\xi$(M)\\,$\\propto$\\,M$^{-\\alpha}$\\,), and then turns over at $\\sim$0.1\\,M$_\\odot$. This rise agrees with the mass function inferred by previous studies, including a survey based on proper motion and photometry. In contrast, the mass function differs significantly from that measured for the Hyades, an open cluster with a similar age ($\\tau$\\,$\\sim$\\,600\\,Myr). Possible reasons are that the clusters did not have the same initial mass function, or that dynamical evolution (e.g.\\ evaporation of low mass members) has proceeded differently in the two clusters. Although different binary fractions could cause the observed (i.e.\\ system) mass functions to differ, there is no evidence for differing binary fractions from measurements published in the literature. Of our cluster candidates, six have masses predicted to be equal to or below the stellar/substellar boundary at 0.072\\,M$_\\odot$.} {} ", "introduction": "Several publications in the past decade have been concerned with the mass function (MF) of low mass stellar and substellar populations in open clusters, including $\\sigma$~Orionis (\\citealt{bejar2002}, \\citealt{caballero2007}), the Orion Nebula Cluster (\\citealt{hillenbrand2000}, \\citealt{slesnick2004}), IC~2391 (\\citealt{barrado2004}, \\citealt{boudreault2009}), the Pleiades (\\citealt{moraux2003}, \\citealt{lodieu2007}), and the Hyades (\\citealt{reid99}, \\citealt{bouvier2008}), to name just a few. Studies of relatively old open clusters (age\\,$\\gtrsim$\\,100\\,Myr) are important for the following two reasons in particular. First, they permit a study of the intrinsic evolution of brown dwarfs (BDs), e.g.\\ their luminosity and effective temperature, which constrains structural and atmospheric models. Second, together with younger clusters we can investigate how BD populations as a whole evolve and thus probe the efficiency with which BDs evaporate from clusters to populate the Galactic field. Numerical simulations of cluster evolution have demonstrated that the MFs can evolve through dynamical interaction (\\citealt{marcos2000}; \\citealt{adams2002b}). These interactions result in a decrease of the open cluster BD (and low-mass star) population. This has been observed by \\cite{bouvier2008} from a comparison of the Pleiades (120\\,Myr) and Hyades (625\\,Myr) mass functions. Many earlier studies of the substellar MF have focused on young open clusters with ages less than $\\sim$\\,100\\,Myr, and in many cases much younger ($< 10$\\,Myr). This is partly because BDs are bright when they are young (lacking a significant nuclear energy source, they cool as they age), thus easing detection of the least massive objects. However, youth presents difficulties. First, intra-cluster extinction plagues the determination of the intrinsic luminosity function from the measured photometry. Second, at these ages the BD models have large(r) uncertainties (\\citealt{baraffe2002}). Estimates of the substellar MF in very young clusters (age $\\lesssim$1\\,Myr) might be unreliable due to these modelling uncertainties (\\citealt{chabrier2005}). BDs in older clusters suffer less from these problems, but have the disadvantage that much deeper surveys are required to detect them. The old open cluster Praesepe is an interesting target considering its age and distance. It is located at a distance of 190$^{+6.0}_{-5.8}$\\,pc (based on parallax measurements from the new Hipparcos data reduction, \\citealt{Leeuwen2009}) and has an age of 590$^{+150}_{-120}$\\,Myr (by isochrone fitting in the Hertzsprung-Russell diagram; \\citealt{fossati2008}). The extinction towards this cluster is low, $E(B-V)$\\,=\\,0.027$\\pm$0.004\\,mag (\\citealt{taylor2006}), while determinations of the metallicity of Praesepe yield some discrepancies: [Fe/H]\\,=\\,0.038$\\pm$0.039, \\cite{friel1992}; +0.13$\\pm$0.10, \\cite{boesgaard1988}; 0.11$\\pm$0.03 from spectroscopy and 0.20$\\pm$0.04 from photometry, \\cite{an2007}; +0.27$\\pm$0.10, \\cite{pace2008}. \\cite{hambly1995} presented a $\\sim$19\\,deg$^2$ survey of the Praesepe cluster down to masses of $\\sim$0.1\\,M$_\\odot$ and observed a rise of the MF at the lowest masses. They concluded that this implied a large population of BDs. A shallow survey complete to $I$\\,=\\,21.2\\,mag, $R$\\,=\\,22.2\\,mag over 800\\,arcmin$^2$ uncovered one spectrally confirmed very low-mass star or BD (spectral type of M8.5V) with a model-dependent mass of 0.063--0.084\\,M$_\\odot$ (\\citealt{Ma_98}). A survey over the central 1\\,deg$^2$ with 10$\\sigma$ limits of $R$\\,=\\,21.5, $I$\\,=\\,20.0 and $Z$\\,=\\,21.5\\,mag revealed 19~BD candidates and the first MF determination of Praesepe down to the substellar limit, but without spectral confirmation (\\citealt{pinfield97}). Subsequent infrared photometry of the sample reduced this number to nine candidates (\\citealt{hodgkin99}). \\cite{adams2002a} presented a 100\\,deg$^2$ study of Praesepe using 2MASS (Two-Micron All Sky Survey) data and Palomar survey photographic plates, from which they derived proper motions. They determined the radial profile of this cluster but their MF does not reach the substellar regime. A more recent proper motion survey of Praesepe covers a much larger area (300\\,deg$^2$; \\citealt{kraus2007}), but does not reach the BD regime either (the limit is $\\sim$0.12\\,M$_\\odot$). Finally, the most recent substellar MF determination of Praesepe was published by \\cite{gonzalez-garcia2006} and extends to a 5$\\sigma$ detection limit of $i=24.5$\\,mag corresponding to 0.050--0.055\\,M$_\\odot$. They identified one new substellar candidate, but their survey covers only 1177 arcmin$^2$. In this paper, we present the results of a program to study, in detail, the MF of Praesepe down to the substellar regime. Our photometric survey is, as with \\cite{gonzalez-garcia2006}, the deepest so far in optical and near-infrared (NIR) bands, with 5$\\sigma$ detection limits of $I_{\\rm c}=23.4$ and $J=20.0$ (corresponding to a mass limit of about 0.05\\,M$_\\odot$), but covers more than nine times the area. Our paper is structured as follows. We first present the data set, reduction procedure and calibration in section \\ref{obs-data-calib}. We then discuss our candidate selection procedure in section \\ref{selection} and the survey results in section \\ref{results-survey} before discussing the derived MF in section \\ref{mf-variation}. We conclude in section \\ref{conclusions}. ", "conclusions": "Conclusions} We have presented the results of a survey to study the mass function of the old open cluster Praesepe. The survey consisted of optical $I_{\\rm c}$-band photometry and NIR $J$ and $K_{\\rm s}$-band photometry with a total coverage of 3.1\\,deg$^2$, down to the substellar regime, with a 5$\\sigma$ detection limit corresponding to 0.05\\,M$_\\odot$ (the detection completeness to this level is $\\sim$87\\%). Our final sample yields 123 photometric cluster member candidates based on a selection assuming a dust-free atmosphere and 27 photometric cluster candidates based on a selection assuming a dusty atmosphere. We estimate the contamination by field M-dwarfs to be 13\\% or less. Among our cluster candidates, six objects have theoretical masses equal to or less than the stellar/substellar boundary at 0.072\\,M$_\\odot$. We observed that the MF of Praesepe is characterized by a rise in the number of objects from 0.6\\,M$_\\odot$ down to 0.1\\,M$_\\odot$, followed by a turn-over in the MF at $\\sim$0.1\\,M$_\\odot$. The rise is in agreement with the Praesepe MFs derived in several previous studies (\\citealt{hambly1995}; \\citealt{kraus2007}; \\citealt{baker2009}) but disagrees with \\cite{adams2002a}. We have compared the mass function of Praesepe with one derived for the Hyades and have observed a significant difference: while the Hyades has a maximum at 0.35\\,M$_\\odot$, Praesepe has a maximum at a much lower mass, 0.1\\,M$_\\odot$. Assuming that they have similar ages (as main sequence fitting suggests), we conclude that the clusters either had different {\\em initial} mass functions or that dynamical interaction has modified the evolution of one or both. More specifically, in the latter case, dynamical evaporation does not seem to have influenced the Hyades and Praesepe in the same way. A difference in the binary fraction or mass ratios could also cause a difference in the mass functions, but determinations of these are not yet precise enough to suggest any difference." }, "0910/0910.4659_arXiv.txt": { "abstract": "The BEST wide-angle telescope installed at the Observatoire de Haute-Provence and operated in remote control from Berlin by the Institut f\\\"ur Planetenforschung, DLR, has observed the CoRoT target fields prior to the mission. The resulting archive of stellar photometric lightcurves is used to search for deep transit events announced during CoRoT's alarm-mode to aid in fast photometric confirmation of these events. The \"initial run\" field of CoRoT (IRa01) has been observed with BEST in November and December 2006 for 12 nights. The first \"long run\" field (LRc01) was observed from June to September 2005 for 35 nights. After standard CCD data reduction, aperture photometry has been performed using the ISIS image subtraction method. About 30,000 lightcurves were obtained in each field. Transits of the first detected planets by the CoRoT mission, CoRoT-1b and CoRoT-2b, were found in archived data of the BEST survey and their lightcurves are presented here. Such detections provide useful information at the early stage of the organization of follow-up observations of satellite alarm-mode planet candidates. In addition, no period change was found over $\\sim$4 years between the first BEST observation and last available transit observations. ", "introduction": "The Berlin Exoplanet Search Telescope (BEST) is a 19.5 cm aperture wide angle telescope dedicated to time-series photometric observations \\cite{Rauer2004}. The main purpose of the instrument is to provide ground-based support to the CoRoT space mission (CNES; \\cite{Baglin2006}). The mission target fields are observed at least one year prior to the observations of the spacecraft. Therefore, planetary transit candidates found by CoRoT at bright stars can be searched for quickly in the BEST archive to confirm the transit event and to check the ephemeris. In addition, BEST data sets are used to search for new variable stars in the CoRoT target fields which are input to the additional science program, like e.g. eclipsing binary stars (\\cite{Karoff2007}, \\cite{Kabath07}, \\cite{Kabath08}). Furthermore, information on variable stars nearby CoRoT transit candidates can be used to help disentangling crowding problems during the CoRoT lightcurve analysis. The transit signals of the first two planets detected by CoRoT (CoRoT-1b and CoRoT-2b) (\\cite{barge2008}, \\cite{alonso2008}), were found using the \"alarm detection mode\" of the mission, during ongoing observations of the respective target fields. Since both planets orbit relatively bright stars, we searched for signatures of the transit candidates in our BEST data archive. Partial transit events of the planets were recorded by BEST in December 2006 and summer 2005 during observations of the CoRoT \"initial run\" and first \"long run\" fields. In the following, we describe these pre-discovery observations of CoRoT-1b and CoRoT-2b. ", "conclusions": "We present pre-discovery observations of the first two planets detected by the CoRoT space mission, CoRoT-1b and CoRoT-2b. The transit events were detected in the BEST data archive, based on the epochs determined from the CoRoT data (\\cite{barge2008}, \\cite{alonso2008}). Although the observational duty cycle was severely affected by bad weather conditions, partial transits of both planets were detected by BEST. A transit of CoRoT-1b was observed on 10 December 2006. Transits of CoRoT-2b were observed in 14 July 2005, 28 July 2005 and 04 Aug 2005. No significant O-C deviation in comparison to the ephemerides of \\cite{alonso2008} was found. When planetary candidates are first announced from the CoRoT alarm-mode, their ephemeris can be based on few transit events only. At this point, confirmation from ground-based observations taken one or more years prior to the spacecraft observations aids in checking the ephemeris quickly. In addition, these observations confirm the transit on the prime target and help identifying close possibly contaminating variable stars. We will continue to use the BEST data archive for this purpose in future to support the planet search in particular during the alarm-mode of CoRoT, when primarily deep transits around bright stars are detected which overlap with the detection range of BEST." }, "0910/0910.2887_arXiv.txt": { "abstract": "{The findings of more than 350 extrasolar planets, most of them nontransiting Hot Jupiters, have revealed correlations between the metallicity of the main-sequence (MS) host stars and planetary incidence. This connection can be used to calculate the planet formation probability around other stars, not yet known to have planetary companions. Numerous wide-field surveys have recently been initiated, aiming at the transit detection of extrasolar planets in front of their host stars. Depending on instrumental properties and the planetary distribution probability, the promising transit locations on the celestial plane will differ among these surveys.} {We want to locate the promising spots for transit surveys on the celestial plane and strive for absolute values of the expected number of transits in general. Our study will also clarify the impact of instrumental properties such as pixel size, field of view (FOV), and magnitude range on the detection probability.} {We used data of the Tycho catalog for $\\approx~1$ million objects to locate all the stars with $0^\\mathrm{m}~\\lesssim~m_\\mathrm{V}~\\lesssim~11.5^\\mathrm{m}$ on the celestial plane. We took several empirical relations between the parameters listed in the Tycho catalog, such as distance to Earth, $m_\\mathrm{V}$, and $(B-V)$, and those parameters needed to account for the probability of a star to host an observable, transiting exoplanet. The empirical relations between stellar metallicity and planet occurrence combined with geometrical considerations were used to yield transit probabilities for the MS stars in the Tycho catalog. Magnitude variations in the FOV were simulated to test whether this fluctuations would be detected by BEST, XO, SuperWASP and HATNet.} {We present a sky map of the expected number of Hot Jupiter transit events on the basis of the Tycho catalog. Conditioned by the accumulation of stars towards the galactic plane, the zone of the highest number of transits follows the same trace, interrupted by spots of very low and high expectation values. The comparison between the considered transit surveys yields significantly differing maps of the expected transit detections. While BEST provides an unpromising map, those for XO, SuperWASP, and HATNet show FsOV with up to 10 and more expected detections. The sky-integrated magnitude distribution predicts 20 Hot Jupiter transits with orbital periods between 1.5\\,d and 50\\,d and $m_\\mathrm{V}~<~8^\\mathrm{m}$, of which two are currently known. In total, we expect 3412 Hot Jupiter transits to occur in front of MS stars within the given magnitude range. The most promising observing site on Earth is at latitude $~=~-1$.} {} ", "introduction": "\\label{sec:intro} A short essay by Otto Struve \\citep{1952Obs....72..199S} provided the first published proposal of transit events as a means of exoplanetary detection and exploration. Calculations for transit detection probabilities \\citep{1971Icar...14...71R, 1984Icar...58..121B, 2006AcA....56..183P} and for the expected properties of the discovered planets have been done subsequently by many others \\citep{2005A&A...442..731G, 2007A&A...475..729F, 2008ApJ...686.1302B}. Until the end of the 1990s, when the sample of known exoplanets had grown to more than two dozen \\citep{2000ApJ...532L..51C}, the family of so-called `Hot Jupiters', with 51 Pegasi as their prototype, was unknown and previous considerations had been based on systems similar to the solar system. Using geometrical considerations, \\citet{1971Icar...14...71R}\\footnote{A correction to his Eq. (2) is given in \\citet{1984Icar...58..121B}.} found that the main contribution to the transit probability of a solar system planet would come from the inner rocky planets. However, the transits of these relatively tiny objects remain undetectable around other stars as yet. The first transit of an exoplanet was finally detected around the sun-like star HD209458 \\citep{2000ApJ...529L..45C, 2000A&A...359L..13Q}. Thanks to the increasing number of exoplanet search programs, such as the ground-based Optical Gravitational Lensing Experiment (OGLE) \\citep{1992AcA....42..253U}, the Hungarian Automated Telescope (HAT) \\citep{2002PASP..114..974B, 2004PASP..116..266B}, the Super Wide Angle Search for Planets (SuperWASP) \\citep{2003ASPC..294..405S}, the Berlin Exoplanet Search Telescope (BEST) \\citep{2004PASP..116...38R}, XO \\citep{2005PASP..117..783M}, the Transatlantic Exoplanet Survey (TrES) \\citep{2007ASPC..366...13A}, and the Tautenburg Exoplanet Search Telescope (TEST) \\citep{2009IAUS..253..340E} and the space-based missions `Convection, Rotation \\& Planetary Transits' (CoRoT) \\citep{2002ESASP.485...17B} and Kepler \\citep{2007CoAst.150..350C}, the number of exoplanet transits has grown to 62 until September $1^\\mathrm{st}$ 2009\\footnote{Extrasolar Planets Encyclopedia (EPE): www.exoplanet.eu. Four of these 62 announced transiting planets have no published position.} and will grow drastically within the next years. These transiting planets have very short periods, typically $<~10$\\,d, and very small semimajor axes of usually $<~0.1$\\,AU, which is a selection effect based on geometry and Kepler's third law \\citep{1619QB41.K38.......}. Transiting planets with longer periods present more of a challenge, since their occultations are less likely in terms of geometrical considerations and they occur less frequently. Usually, authors of studies on the expected yield of transit surveys generate a fictive stellar distribution based on stellar population models. \\citet{2007A&A...475..729F} use a Monte-Carlo procedure to synthesize a fictive stellar field for OGLE based on star counts from \\citet{2006AcA....56....1G}, a stellar metallicity distribution from \\citet{2004A&A...418..989N}, and a synthetic structure and evolution model of \\citet{2003A&A...409..523R}. The metallicity correlation, however, turned out to underestimate the true stellar metallicity by about 0.1\\.dex, as found by \\citet{2004A&A...415.1153S} and \\citet{2005ApJ...622.1102F}. In their latest study, \\citet{2009A&A...504..605F} first generate a stellar population based on the Besan\\c{c}on catalog from \\citet{2003A&A...409..523R} and statistics for multiple systems from \\citet{1991A&A...248..485D} to apply then the metallicity distribution from \\citet{2004A&A...415.1153S} and issues of detectability \\citep{2006MNRAS.373..231P}. \\citet{2008ApJ...686.1302B} rely on a Galactic structure model by \\citet{1980ApJS...44...73B}, a mass function as suggested by \\citet{2002AJ....124.2721R} based on Hipparcos data, and a model for interstellar extinction to estimate the overall output of the current transit surveys TrES, XO, and Kepler. In their paper on the number of expected planetary transits to be detected by the upcoming Pan-STARRS survey \\citep{2004SPIE.5489...11K}, \\citet{2009A&A...494..707K} also used a Besan\\c{c}on model as presented in \\citet{2003A&A...409..523R} to derive a brightness distribution of stars in the target field and performed Monte-Carlo simulations to simulate the occurrence and detections of transits. These studies include detailed observational constraints such as observing schedule, weather conditions, and exposure time and issues of data reduction, e.g. red noise and the impact of the instrument's point spread function. In our study, we rely on the extensive data reservoir of the Tycho catalog instead of assuming a stellar distribution or a Galactic model. We first estimate the number of expected exoplanet transit events as a projection on the complete celestial plane. We refer to recent results of transit surveys such as statistical, empirical relationships between stellar properties and planetary formation rates. We then use basic characteristics of current low-budget but high-efficiency transit programs (BEST, XO, SuperWASP, and HATNet), regardless of observational constraints mentioned above, and a simple model to test putative transits with the given instruments. With this procedure, we yield sky maps, which display the number of expected exoplanet transit detections for the given surveys, i.e. the transit sky as it is seen through the eyeglasses of the surveys. The Tycho catalog comprises observations of roughly 1 million stars taken with the Hipparcos satellite between 1989 and 1993 \\citep{1997yCat.1239....0E, 1997ESASP.402...25H}. During the survey, roughly 100 observations were taken per object. From the derived astrometric and photometric parameters, we use the right ascension ($\\alpha$), declination ($\\delta$), the color index $(B-V)$, the apparent visible magnitude $m_\\mathrm{V}$, and the stellar distance $d$ that have been calculated from the measured parallax. The catalog is almost complete for the magnitude limit $m_\\mathrm{V}~\\lesssim~11.5^\\mathrm{m}$, but we also find some fainter stars in the list. ", "conclusions": "\\label{sec:discussion} Our values for the XO project are much higher than those provided by \\citet{2008ApJ...686.1302B}, who also simulated the expected exoplanet transit detections of XO. This is due to their much more elaborate inclusion of observational constraints such as observational cadence, i.e. hours of observing per night, meteorologic conditions, exposure time, and their approach of making assumptions about stellar densities and the Galactic structure instead of using catalog-based data as we did. Given these differences between their approach and ours, the results are not one-to-one comparable. While the study of \\citet{2008ApJ...686.1302B} definitely yields more realistic values for the expected number of transit detections considering all possible given conditions, we provide estimates for the celestial distribution of these detections, neglecting observational aspects. In addition to the crucial respects that make up the efficiency of the projects, as presented in Table \\ref{tab:surveys}, SuperWASP and HATNet benefit from the combination of two observation sites and several cameras, while XO also takes advantage of twin lenses but a single location. Each survey uses a single camera type and both types have similar properties, as far as our study is concerned. The transit detection maps in Fig. \\ref{fig:surveys} refer to a single camera of the respective survey. The alliance of multiple cameras and the diverse observing strategies among the surveys \\citep{2005PASP..117..783M, 2009IAUS..253...29C} bias the speed and efficiency of the mapping procedure. This contributes to the dominance of SuperWASP (18 detections, 14 of which have published positions)\\footnote{EPE as of September $1^\\mathrm{st}$ 2009} over HATNet (13 detections, all of which have published positions)\\footnotemark[3], XO (5 detections, all of which have published positions)\\footnotemark[3], and BEST (no detection)\\footnotemark[3]. It is inevitable that a significant fraction of unresolved binary stars within the Tycho data blurs our results. The impact of unresolved binaries without physical interaction, which merely happen to be aligned along the line of sight, is significant only in the case of extreme crowding. As shown by \\citet{2007ASPC..366..283G}, the fraction of planets not detected because of blends is typically lower than 10\\,\\%. The influence of unresolved physical binaries will be higher. Based on the empirical period distribution for binary stars from \\citet{1991A&A...248..485D}, \\citet{2008ApJ...686.1302B} estimate the fraction of transiting planets that would be detected despite the presence of binary systems to be $\\approx~70$\\,\\%. Both the contribution of binary stars aligned by chance and physically interacting binaries result in an overestimation of our computations of $\\approx~40$\\,\\%, which is of the same order as uncertainties arising from the empirical relationships we use. Moreover, as \\citet{2006MNRAS.367.1103W} have shown, the density of eclipsing stellar binary systems increases dramatically towards the Galactic center. To control the fraction of false alarms, efficient data reduction pipelines, and in particular data analysis algorithms, are necessary \\citep{2006MNRAS.365..165S}. Recent evidence for the existence of ultra-short period planets around low-mass stars \\citep{2009IAUS..253...45S}, with orbital periods $~<~1$\\,d, suggests that we underestimated the number of expected transits to occur, as presented in Sect. \\ref{sub:occurance}. The possible underestimation of exoplanets occurring at $\\mathrm{[Fe/H]}_\\star~<~0$ also contributes to a higher number of transits and detections than we computed here. Together with the fact that the Tycho catalog is only complete to $m_\\mathrm{V} \\lesssim 11.5^\\mathrm{m}$, whereas the surveys considered here are sensitive to slightly fainter stars (see Table \\ref{tab:surveys}), these trends towards higher numbers of expected transit detections might outweigh the opposite effect of unresolved binary stars. A radical refinement of both our maps for transits occurrence and detections will be available within the next few years, once the `Panoramic Survey Telescope and Rapid Response System' (Pan-STARRS) \\citep{2002SPIE.4836..154K} will run to its full extent. Imaging roughly 6000 square degrees every night with a sensitivity down to $m_\\mathrm{V} \\approx 24$, this survey will not only drastically increase the number of cataloged stars -- thus enhance our knowledge of the localization of putative exoplanetary transits -- but could potentially detect transits itself \\citep{2009ApJ...704.1519D}. The Pan-STARRS catalog will provide the ideal sky map, on top of which an analysis presented in this paper can be repeated for any ground-based survey with the aim of localizing the most appropriate transit spots on the celestial plane. The bottleneck for the verification of transiting planets, however, is not the localization of the most promising spots but the selection of follow-up targets accessible with spectroscopic instruments. The advance to fainter and fainter objects thus won't necessarily lead to more transit confirmations. Upcoming spectrographs, such as the ESPRESSO{\\MVAt}VLT and the CODEX{\\MVAt}E-ELT \\citep{2008PhST..130a4007P}, can be used to confirm transits around fainter objects. These next-generation spectrographs that will reveal Doppler fluctuations on the order of cm$\\cdot$s$^{-1}$ will also enhance our knowledge about Hot Neptunes and Super-Earths, which the recently discovered transits of GJ\\,436\\,b \\citep{2004ApJ...617..580B}, HAT-P-11\\,b \\citep{2009arXiv0901.0282B}, and CoRoT-7b \\citep{2009arXiv0908.0241L} and results from \\citet{2009IAUS..253..502L} predict to be numerous. Further improvement of our strategy will emerge from the findings of more exoplanets around MS stars and from the usage of public data reservoirs like the NASA Star and Exoplanet Database\\footnote{http://nsted.ipac.caltech.edu}, making assumptions about the metallicity distribution of planet host stars and the orbital period distribution of exoplanets more robust." }, "0910/0910.0882_arXiv.txt": { "abstract": "\\noindent Globular clusters are an important test bed for Newtonian gravity in the weak-acceleration regime, which is vital to our understanding of the nature of the gravitational interaction. Recent claims have been made that the velocity dispersion profiles of globular clusters flatten out at large radii, despite an apparent paucity of dark matter in such objects, indicating the need for a modification of gravitational theories. We continue our investigation of this claim, with the largest spectral samples ever obtained of 47 Tucanae and M55. Furthermore, this large sample allows for an accurate metallicity calibration based on the equivalent widths of the calcium triplet lines and $K$ band magnitude of the Tip of the Red Giant Branch. Assuming an isothermal distribution, the rotations of each cluster are also measured with both clusters exhibiting clear rotation signatures. The global velocity dispersions of NGC 121 and Kron 3, two globular clusters in the Small Magellanic Cloud, are also calculated. By applying a simple dynamical model to the velocity dispersion profiles of 47 Tuc and M55, we calculate their mass-to-light profiles, total masses and central velocity dispersions. We find no statistically significant flattening of the velocity dispersion at large radii for M55, and a marked {\\it increase} in the profile of 47 Tuc for radii greater than approximately half the tidal radius. We interpret this increase as an evaporation signature, indicating that 47 Tuc is undergoing, or has undergone, core-collapse, but find no requirement for dark matter or a modification of gravitational theories in either cluster. ", "introduction": "The nature of the gravitational interaction is one of the most important concepts in astrophysics, yet complete comprehension of this interaction is still elusive. The so-called Pioneer and Fly-by anomalies, where spacecraft exhibit behaviour that is unexpected from Newtonian and general relativity gravitation theories, outline this lack of understanding \\cite[see][and references therein]{Anderson02,deDiego08}, although these examples may have more mundane explanations. More importantly, it has been claimed that several globular clusters (GCs; $\\omega$~Centauri, M15, M30, M92 and M107) exhibit a flattening of their velocity dispersion profiles at radii $R\\sim\\frac{r_t}{2}$, where $r_t$ is the tidal radius of the cluster \\citep{Scarpa03,Scarpa04a,Scarpa04b}. The authors claim that either dark matter (DM), or a modification of gravitational theory, is required to explain their results. Modified theories of gravity \\cite[MOG; see][for a review of modified gravity theories]{Durrer08} and those of Newtonian dynamics \\citep[MOND;][]{Milgrom83} have been shown to solve some of the discrepancies. However, these are not universal theories and to-date have only been applied to specific instances \\cite[e.g. the Bullet Cluster and galaxy rotation curves; see][respectively]{Angus06,Sanders07}. MOG theories diverge from Newtonian gravity in the high-acceleration regime and MOND diverges from Newton in the low-acceleration regime. Therefore, if either of these theories were correct, the effect should be measurable at the predicted accelerations. Independent of MOG or MOND theories, testing the gravitational interaction at low accelerations is essential to the overall understanding of gravity. Globular clusters are an ideal testing ground for weak-field gravitation because the accelerations experienced by stars at large radii are below the limit where DM, or a modified gravitation theory, is required to explain observations in many dynamical systems \\cite[$a_0\\approx1.2\\times10^{-10}$\\,m\\,s$^{-2}$;][]{Scarpa07}. Furthermore, they are thought to contain little, or no, dark matter -- indicated by dynamical models \\citep{Phinney93}, {\\it N}-body simulations \\citep{Moore96}, observations of GC tidal tails \\citep{Odenkirchen01}, dynamical and luminous masses of GCs \\citep{Mandushev91} and the lack of microlensing events from GC-mass dark haloes \\citep{Navarro97,Ibata02}, although this is still under debate. GCs are also located at varying distances from the centre of the Galaxy, so that if all exhibit similar behaviour, Galactic influences cannot be the primary cause. In \\citealt{Lane09} (hereafter \\citetalias{Lane09}) we calculated the velocity dispersions and mass-to-light profiles of M22, M30, M53 and M68. Our conclusions were that there is no requirement for significant quantities of dark matter, or a modification of Newtonian gravity, to explain the kinematics of any of these clusters. In the current paper we continue this investigation with the largest spectroscopic dataset to date of 47 Tucanae and M55. We begin by describing the data acquisition/reduction (Section \\ref{data}) and the membership selection for each cluster (Section \\ref{membership}). Our large samples allow us to determine a metallicity calibration based on the Tip of the Red Giant Branch (TRGB; Section \\ref{metaldist}), as well as the rotations (Section \\ref{rotation}), systemic velocities and velocity dispersions (Section \\ref{dispersions}), and mass-to-light profiles (Section \\ref{ML}) -- where we also place limits on the DM content of each cluster from their velocity dispersions and mass-to-light profiles. Finally, our concluding remarks are presented in Section \\ref{conclusions}. ", "conclusions": "Having the largest sample of spectra ever obtained for 47 Tuc and M55, we were able to produce a very accurate calibration of [Fe/H] based on the equivalent width of the calcium triplet lines and the $K$ band magnitude of the TRGB. This method is similar to that by \\cite{Cole04} and \\cite{Warren09}, except that we use the TRGB instead of the HB which means this method can be used for much more distant objects. We calculated the rotation of our clusters assuming them to be isothermal. The rotation of 47 Tuc is $\\sim2.2\\pm0.2$ with an approximate projected rotational axis along the line PA = $40^\\circ-220^\\circ$, and M55 exhibits rotation at a level of $0.25\\pm0.09$\\,km\\,s$^{-1}$ and has an approximate axis of rotation along the line PA = $65^\\circ-245^\\circ$. For 47 Tuc, the rotation amplitude is in good agreement with previous work \\cite[e.g.][]{Meylan86}. The only previous study estimating the rotation of M55 \\citep{Szekely07} found a value about twice that of the current work, but \\cite{Szekely07} had a sample size approximately half that of ours, which may explain this discrepancy. Our calculated velocity dispersion profiles of 47 Tuc and M55 provide no evidence that either DM or a modification of current gravitational theories are required to reconcile their kinematic properties, corroborating previous work in \\citetalias{Lane09}. The dynamics of M55 are well described by a purely analytic \\citet{Plummer11} model, which indicates that Newtonian gravity adequately describes its velocity dispersions, and shows no breakdown of Newtonian gravity at $a_0\\approx1.2\\times10^{-10}$\\,ms$^{-2}$, as has been claimed for other GCs. The internal dynamics of 47 Tuc (for $Rr_t/2$, exactly the region where evaporation due to two-body interactions in the core should be observable, especially for GCs undergoing core-collapse or in a post-core-collapse state. We interpret this increase in velocity dispersion as evaporation, and hence that this cluster is either presently in a state of core-collapse, or is a post-core-collapse GC. This adds to the growing evidence that 47 Tuc is currently undergoing a dynamical phase change \\cite[e.g.][and references therein]{Gebhardt95b,Robinson95,Howell00}. A full analysis of the outer regions of this apparently evaporating cluster will be performed in a subsequent paper. We used a Plummer model to determine the total mass, scale radius ($r_s$), and the M/L$_{\\rm V}$ profile for each cluster. We find that neither cluster has M/L$_{\\rm V}\\gg1$, indicating that DM is not dominant. Within the uncertainties, our estimated cluster masses match those in the literature well, as do the M/L$_{\\rm V}$ ratios \\cite[e.g.][]{Meylan89,Pryor93,Meziane96,Kruijssen09}. The mass-to-light profiles produced by \\cite{Gebhardt95a} cannot be compared to the current work because they sampled the inner $10'$ for which the mass is uncertain. Note that we consider using mass-to-light {\\it profiles} is a more accurate method for calculating M/L$_{\\rm V}$ than using the core mass and surface brightness, because of this uncertainty. While our results strongly indicate that the current understanding of globular clusters being dark matter poor, and their dynamics explained by standard Newtonian gravity, more robust dynamical modelling is required for confirmation." }, "0910/0910.5469_arXiv.txt": { "abstract": "We analyze the statistics and star formation rate obtained in high-resolution numerical experiments of forced supersonic turbulence, and compare with observations. We concentrate on a systematic comparison of solenoidal (divergence-free) and compressive (curl-free) forcing (Federrath et~al.~2009~a,b), which are two limiting cases of turbulence driving. Our results show that for the same RMS Mach number, compressive forcing produces a three times larger standard deviation of the density probability distribution. When self-gravity is included in the models, the star formation rate is more than one order of magnitude higher for compressive forcing than for solenoidal forcing. ", "introduction": " ", "conclusions": "" }, "0910/0910.0285_arXiv.txt": { "abstract": "% In this paper we discuss recent applications of the Smoothed Particle Hydrodynamics (SPH) method to the simulation of supersonic turbulence in the interstellar medium, as well as giving an update on recent algorithmic developments in solving the equations of magnetohydrodynamics (MHD) in SPH. Using high resolution calculations (up to 134 million particles), we find excellent agreement with grid-based results on a range of measures including the power spectrum slope in both the velocity field and the density-weighted velocity $\\rho^{1/3} v$, the latter showing a Kolmogorov-like $k^{-5/3}$ scaling as proposed by Kritsuk et al. (2007). We also find good agreement on the statistics of the Probability Distribution Function (PDF) and structure functions, independently confirming the scaling found by \\citet*{sfk08}. On Smoothed Particle Magnetohydrodynamics (SPMHD) we have recently wasted a great deal of time and effort investigating the vector potential as an alternative to the Euler potentials formulation, in the end concluding that using the vector potential has even more severe problems than the standard (${\\bf B}$-field based) SPMHD approach. ", "introduction": "Turbulence and magnetic fields are thought to be two of the most important ingredients in the star formation process, so much so that there remains ongoing debate --- both observational and theoretical --- as to which one is \\emph{the} controlling factor \\citep[e.g.][]{mk04,km09}. We have only recently begun to understand the role of either in depth, primarily as a result of our increased ability to simulate both in numerical calculations of the star formation process. Several authors have proposed that turbulence in the interstellar medium (ISM) can be used to predict the form of the stellar initial mass function (IMF) thus providing a `theory' of star formation \\citep{pn02,HennebelleChabrier2008}. Any such theory requires consensus on the basic statistical characteristics of turbulence in the supersonic regime and assumes that these are universal --- that is, independent of boundary conditions and driving mechanisms --- for which there is some observational support \\citep[e.g.][]{hb04}. However there exists considerable disagreement between results obtained using different numerical codes, most recently between \\citet{padoanetal07} who claim that the statistics of turbulence are universal and \\citet{bpetal06} who claim that they are not, based on calculations utilising both a Smoothed Particle Hydrodynamics (SPH) code and a grid-based Total-Variation-Diminishing (TVD) code. This kind of disagreement, and the need within star formation theory for a consensus on the basic statistics of supersonic turbulence has prompted at least two major code comparison projects over the last few years, the ``Potsdam'' comparison \\citep{kitsionasetal09}, comparing decaying, hydrodynamic turbulence and the KITP comparison (unpublished), comparing decaying hydrodynamic and magnetohydrodynamic (MHD) turbulence. However these comparisons suffer from the limited time evolution that can be obtained from a decaying turbulence simulation as well as the numerical issues posed by starting from an evolved snapshot produced by a particular code. We have therefore undertaken our own very detailed comparison of \\emph{driven} turbulence using just two codes, an SPH code, \\textsc{phantom}, and a grid code, \\textsc{flash} (used in uniform grid mode) taken to be broadly representative of their class of codes, in order to see whether agreement can be reached (and at what resolution) on the statistics of supersonic turbulence appropriate to the ISM. The results are published in full in \\citet{pf10} and we only provide a snapshot here of the main results (\\S\\ref{sec:turbulence}), referring the reader to that paper for more detailed information. What we are not yet able to address is the combination of \\emph{driven} turbulence and MHD in SPH, primarily because of the limitations posed by the Euler potentials formulation used as the only method that sufficiently maintains the divergence-free ($\\nabla\\cdot{\\bf B} = 0$) constraint on the magnetic field such that star formation calculations can be performed \\citep[e.g.][]{pb07,pb08,pb09} without the stars themselves exploding due to numerical errors (see \\S\\ref{sec:mhd}). Following a suggestion from Axel Brandenburg, we therefore embarked on a quest to examine whether or not the use of the vector potential could similarly solve the divergence problem in the context of Smoothed Particle Magnetohydrodynamics (SPMHD) without the associated topological restrictions of the Euler potentials \\citep[see][]{brandenburg10}. In short, this turned out not to be the case \\citep[read the horror that is][]{price10}. However, for the pleasure of the reader we describe the tortuous journey in \\S\\ref{sec:mhd} ", "conclusions": "" }, "0910/0910.2249_arXiv.txt": { "abstract": "We report the detection of $\\gamma$-ray pulsations ($\\ge 0.1$ GeV) from PSR~J2229+6114 and PSR~J1048$-$5832, the latter having been detected as a low-significance pulsar by EGRET. Data in the $\\gamma$-ray band were acquired by the Large Area Telescope aboard the \\textit{Fermi Gamma-ray Space Telescope}, while the radio rotational ephemerides used to fold the $\\gamma$-ray light curves were obtained using the Green Bank Telescope, the Lovell telescope at Jodrell Bank, and the Parkes telescope. The two young radio pulsars, located within the error circles of the previously unidentified EGRET sources 3EG~J1048$-$5840 and 3EG~J2227+6122, present spin-down characteristics similar to the Vela pulsar. PSR~J1048$-$5832 shows two sharp peaks at phases $0.15 \\pm 0.01$ and $0.57 \\pm 0.01$ relative to the radio pulse confirming the EGRET light curve, while PSR~J2229+6114 presents a very broad peak at phase $0.49 \\pm 0.01$. The $\\gamma$-ray spectra above 0.1 GeV of both pulsars are fit with power laws having exponential cutoffs near 3 GeV, leading to integral photon fluxes of $(2.19 \\pm 0.22 \\pm 0.32) \\times 10^{-7}$\\,cm$^{-2}$\\,s$^{-1}$ for PSR~J1048$-$5832 and $(3.77 \\pm 0.22 \\pm 0.44) \\times 10^{-7}$\\,cm$^{-2}$\\,s$^{-1}$ for PSR~J2229+6114. The first uncertainty is statistical and the second is systematic. PSR~J1048$-$5832 is one of two LAT sources which were entangled together as 3EG~J1048$-$5840. These detections add to the growing number of young $\\gamma$-ray pulsars that make up the dominant population of GeV $\\gamma$-ray sources in the Galactic plane. ", "introduction": "The nature of unidentified high-energy $\\gamma$-ray sources in the Galaxy was one of the major unanswered questions at the end of the EGRET era. The third EGRET catalog contained 170 unidentified sources, 74 of which were at Galactic latitude $|b|<10$\\degr \\citep{hartman99,bhatta03}. Rotation-powered pulsars are believed to dominate the Galactic $\\gamma$-ray source population (e.g. \\citet{yadiga95}), but their visibility is linked to their beam patterns. Soon after launch, the {\\it Fermi Gamma-ray Space Telescope} began to unveil many 3EG sources, discovering the radio-quiet pulsar in the CTA1 supernova remnant associated with 3EG~J0010+7309 \\citep{abdo08}, detecting the radio-loud pulsar PSR~J2021+3651 (associated with 3EG~J2021+3716) \\citep{abdo09b}, seen independently by {\\em AGILE} \\citep{halpern08}, as well as the radio pulsar PSR~J1028$-$5819, associated with 3EG~J1027$-$5817 \\citep{abdo09c} and new populations of radio-quiet $\\gamma$-ray pulsars, detectable using blind search techniques \\citep{abdo09f}. For the pulsars detected in $\\gamma$ rays, the bulk of the electromagnetic power output is in high energies. The $\\gamma$-ray emission is thus crucial for understanding the emission mechanism which converts the rotational energy of the neutron star into electromagnetic radiation. The discovery of many new $\\gamma$-ray pulsars will provide strong constraints on the location of the $\\gamma$-ray emitting regions, whether above the polar caps \\citep{daugherty96}, or far from the neutron star in the so-called ``outer gaps\" \\citep{romani96}, or in the intermediate regions like the ``slot gap\" \\citep{muslimov04} having ``two-pole caustic\" geometry \\citep{dyks03}. In this paper we report the \\textit{Fermi} detection of the two pulsars PSR~J1048$-$5832 and PSR~J2229+6114, which have spin characteristics similar to other young pulsars typified by the Vela pulsar. \\citet{kramer03} provide a discussion of ``Vela-like'' pulsars. \\\\ PSR~J1048$-$5832 (B1046$-$58) is located in the Carina region at low Galactic latitude ($l=287.42$\\degr, $b=0.58$\\degr). It was discovered during a 1.4 GHz Parkes survey of the Galactic plane and has a period $P~\\sim$123.7~ms \\citep{johnston92}. It has a spin-down luminosity $\\dot{E}=4 \\pi^{2} I (\\dot{P}/P^{3})$ of $2\\times 10^{36}$\\,erg\\,s$^{-1}$, for a moment of inertia $I$ of $10^{45}$\\,g\\,cm$^{2}$, a surface dipole magnetic field strength of $3.5 \\times 10^{12}$\\,G, and a characteristic age $\\tau_{c} = P/2\\dot{P}$ of 20\\,kyr. High-resolution observations of the coincident $ASCA$ source by the \\textit{Chandra X-ray Observatory} and {\\it XMM-Newton\\/} revealed an asymmetric pulsar wind nebula (PWN) of $\\sim6''\\times11''$, surrounding a point source coincident with the pulsar but so far no X-ray pulsations were detected \\citep{gonzalez06}. PSR~J1048$-$5832 has been proposed as a counterpart of the steady EGRET source 3EG~J1048$-$5840 \\citep{fierro95,pivovaroff00,nolan03}, as suggested by positional coincidence, spectral, and energetic properties. Detailed analysis of the EGRET data found possible $\\gamma$-ray pulsations at $E>$400 MeV, a double-peaked light curve with $\\sim 0.4$ peak phase separation (Fig.\\ref{fig:J1048-5832_LC_multi_energy}, \\citealt{kaspi00}). An HI distance determination for the pulsar yields between 2.5 and 6.6 kpc \\citep{johnston96}. The NE2001 model \\citep{cordes02} assigns a distance of 2.7~kpc based partly on the HI distance determination. For this paper we adopt 3~kpc as the distance to the pulsar.\\\\ PSR J2229+6114 is located at ($l$,$b$) = (106.6\\degr,2.9\\degr) within the error box of the EGRET source 3EG~J2227+6122 \\citep{hartman99}. Detected as a compact X-ray source by \\textit{ROSAT} and \\textit{ASCA} observations of the EGRET error box, it was later discovered to be a radio and X-ray pulsar with a period of $P$ = 51.6 ms \\citep{halpern01b}. The radio pulse profile shows a single sharp peak, while the X-ray light curve at 0.8\\ --\\ 10 keV consists of two peaks, separated by $\\Delta\\phi = 0.5$. {\\em AGILE} recently reported the discovery of $\\gamma$-ray pulsations above 100 MeV \\citep{pellizzoni09}. The pulsar is as young as the Vela pulsar (characteristic age $\\tau_{c} = 10$\\,kyr), as energetic ($\\dot{E} = 2.2 \\times 10^{37}$\\,erg\\,s$^{-1}$), and is evidently the energy source of the ``Boomerang'' arc-shaped PWN G106.65+2.96, suggested to be part of the supernova remnant (SNR) G106.3+2.7 discovered by \\citet{joncas90}. Recently, the PWN has been detected at TeV energies by MILAGRO \\citep{abdo09g}. Studies of the radial velocities of both neutral hydrogen and molecular material place the system at $\\sim$ 800 pc \\citep{kothes01}, while \\citet{halpern01a} suggest a distance of 3 kpc estimated from its X-ray absorption. The pulsar DM, used in conjunction with the NE2001 model, yields a distance of 7.5 kpc\\,, significantly above all other estimates. For this paper we again adopt a distance of 3\\,kpc. ", "conclusions": "Although PSR~J1048$-$5832 and PSR~J2229+6114 are both Vela-like in age and spin characteristics, their light curves and derived emission geometries are quite different. Table \\ref{tab:sumres} summarises the main quantities measured for both pulsars. The double-peaked light curve of PSR J1048$-$5832 is nearly identical to that of Vela, whereas its derived efficiency is a factor of 10 larger than that of Vela \\citep{abdo09a}, adopting $f_{\\Omega} = 1$ and $d = 3$\\,kpc. On the contrary, the $\\gamma$-ray efficiency of J2229+6114 is very similar to that of the Vela pulsar, but the pulsar shows a single, large peak similar to PSR~J0357+32 discovered by searching for pulsations at the positions of bright $\\gamma$-ray sources \\citep{abdo09d}. Note that the efficiency of PSR~J2229+6114 would be smaller at the distance of about 1\\,kpc that some authors suggest. \\begin{table*}[t] \\caption{This table summarises the results of the timing and spectral analysis of PSR~J1048$-$5832 and PSR J2229+6114. Statistical and systematics errors are reported.} \\vspace{0.1cm} \\begin{center} \\begin{tabular}{llcc} \\hline Analysis & Parameters & PSR~J1048$-$5832 & PSR~J2229+6449 \\\\ \\hline \\hline \\textbf{Timing results} & Number of pulsed $\\gamma$ rays & 933 $\\pm$ 93 & 1365 $\\pm$ 97\\\\ & Peak position ($\\phi$)& 0.15 $\\pm$ 0.01 $\\pm$ 0.0001 (P1) & 0.49 $\\pm$ 0.01 $\\pm$ 0.001 \\\\ & & 0.57 $\\pm$ 0.01 $\\pm$ 0.0001 (P2) & \\\\ & Peak separation ($\\Delta$) & 0.42 $\\pm$ 0.01 $\\pm$ 0.0001 & \\\\ & Peak FWHM& 0.06 $\\pm$ 0.01 (P1) & 0.23 $\\pm$ 0.03 \\\\ & & 0.10 $\\pm$ 0.02 (P2) & \\\\ \\hline \\textbf{Spectral results} & $^{a}$F (10$^{-7}$ cm$^{-2}$s$^{-1}$) & 2.19 $\\pm$ 0.22 $\\pm$ 0.32 & 3.77 $\\pm$ 0.22 $\\pm$ 0.44 \\\\ & $^{b}$F$_{E}$ (10$^{-11}$ erg cm$^{-2}$s$^{-1}$) & 19.4 $\\pm$ 1.0 $\\pm$ 3.1 & 23.7 $\\pm$ 0.7 $\\pm$ 2.5 \\\\ & $\\Gamma$& 1.38 $\\pm$ 0.06 $\\pm$ 0.12 & 1.85 $\\pm$ 0.06 $\\pm$ 0.10 \\\\ & $^{c}$E$_{c}$ (GeV)& 2.3$^{+0.3}_{-0.4}$ $\\pm$ 0.3 & 3.6$^{+0.9}_{-0.6}$ $\\pm$ 0.6 \\\\ & L$_{\\gamma}$ (10$^{35}$ erg s$^{-1}$) & 2.1 $f_{\\Omega}$ (d/3kpc)$^{2}$ & 2.6 $f_{\\Omega}$ (d/3kpc)$^{2}$ \\\\ & $\\eta_{\\gamma}$ & 0.10 $f_{\\Omega}$ (d/3kpc)$^{2}$& 0.011 $f_{\\Omega}$ (d/3kpc)$^{2}$ \\\\ \\hline \\end{tabular} \\tablenotetext{a}{Integral photon flux ($E>$0.1 GeV)} \\tablenotetext{b}{Integral energy flux ($E>$0.1 GeV)} \\tablenotetext{c}{Energy of an exponential cut-off to a power-law spectrum with index $\\Gamma$.} \\end{center} \\label{tab:sumres} \\end{table*} With the growing number of detected $\\gamma$-ray pulsars, we are beginning to sample a wider variety of emission and viewing geometries, and pulsar ages. The range of light curve morphologies should allow improved constraints on high-energy emission models and a better understanding of the pulsar magnetospheric structure and acceleration process. For example, while many young pulsars, like J1048$-$5832, show Vela-type light curves, a small number are similar to J2229+6114, with a strong P2 component and a weak or absent P1 \\citep{welte09b}. Mapping the angle range over which P1 is missing, especially when viewing angle constraints are available, can help narrow down the high altitude emission zone. A larger pulsar sample also allows a study of evolution of the $\\gamma$-ray beaming and efficiency with pulsar age:~the pulsars seen with \\textit{Fermi}, not including millisecond pulsars, span ages from 10$^{3}$ to $2 \\times 10^{6}$\\,yr \\citep{abdo09i}, with hopes to extend that to larger $\\tau$ (lower $\\dot{E}$) as observations continue. Analysis of the population of pulsars with interpulses \\citep{welte08b} and radio polarization data \\citep{tauris98} have given hints that the magnetic inclination is larger for young pulsars and decreases with age. Thus we may even probe evolution of magnetic alignment during pulsar spindown." }, "0910/0910.5005_arXiv.txt": { "abstract": "We measure shifts of the acoustic scale due to nonlinear growth and redshift distortions to a high precision using a very large volume of high-force-resolution simulations. We compare results from various sets of simulations that differ in their force, volume, and mass resolution. We find a consistency within $1.5-\\sigma$ for shift values from different simulations and derive shift $\\al(z) -1 = (0.300\\pm 0.015)\\% [D(z)/D(0)]^{2}$ using our fiducial set. We find a strong correlation with a non-unity slope between shifts in real space and in redshift space and a weak correlation between the initial redshift and low redshift. Density-field reconstruction not only removes the mean shifts and reduces errors on the mean, but also tightens the correlations. After reconstruction, we recover a slope of near unity for the correlation between the real and redshift space and restore a strong correlation between the initial and the low redshifts. We derive propagators and mode-coupling terms from our \\Nb\\ simulations and compare with the \\Zel\\ approximation and the shifts measured from the $\\chi^2$ fitting, respectively. We interpret the propagator and the mode-coupling term of a nonlinear density field in the context of an average and a dispersion of its complex Fourier coefficients relative to those of the linear density field; from these two terms, we derive a signal-to-noise ratio of the acoustic peak measurement. We attempt to improve our reconstruction method by implementing 2LPT and iterative operations, but we obtain little improvement. The Fisher matrix estimates of uncertainty in the acoustic scale is tested using $5000\\trihGpc$ of cosmological PM simulations from \\citet{Taka09}. At an expected sample variance level of 1\\%, the agreement between the Fisher matrix estimates based on \\citet{SE07} and the \\Nb\\ results is better than 10 \\%. ", "introduction": "In recent years, attention to baryon acoustic oscillations (BAO) as a dark energy probe has increased unprecedentedly due to its robust nature against systematics, and they are now an essential component of most of the major future dark energy surveys under consideration. BAO originate from the sound waves that propagated through the hot plasma of photons and baryons in the very early Universe. At the epoch of recombination, photons and baryons decouple, and as a result, the sound waves freeze out, leaving a distinctive oscillatory feature in the large-scale structure of the cosmic microwave background \\citep[e.g.,][]{Mil99,deB00,Han00,Lee01,HalDasi,Netter02,Pearson03,BenoitArcheops,BennettWmap,Hinshaw07,Hinshaw08} and the matter density fields in Fourier space \\citep[e.g.,][]{Peebles70,SZ70,Bond84,Holtzman89,HS96,Hu96,EH98,Meiksin99}; In configuration space, the BAO appears as a single spherical peak at its characteristic scale. The characteristic physical scale of this oscillatory feature, BAO, is the distance that the sound waves have traveled before the epoch of recombination, which is known as ``sound horizon scale''. This sound horizon scale is and will be measured precisely from current and future CMB data. With the knowledge of the physical scale, BAO can be used as a standard ruler to measure angular diameter distance and Hubble parameter at various redshifts and therefore provide critical information to identify dark energy \\citep[e.g.,][]{Hu96,Eisen03,Blake03,Linder03,Hu03,SE03}. Recently, BAO have been detected from large-scale structure of galaxy distributions and have been used to place an important constraint on dark energy \\citep[][]{Eisen05,Cole05,Hutsi06,Tegmark06,Percival07a,Percival07b,Blake07,Pad07,Okumura08,Estra08,Gazt08a,Gazt08b,Sanchez09,Percival09,Kazin09}. Due to nonlinear structure growth at late times, the oscillatory feature of the BAO is increasingly damped, proceeding from small scales to larger scales, with decreasing redshift \\citep[e.g.,][]{Meiksin99,SE05,Jeong06,ESW07,Crocce08,Mat08}. Redshift distortions enhance a nonlinear damping along the line of sight direction\\citep[e.g.,][]{Meiksin99,SE05,ESW07,Mat08}. Despite the resulting loss of the BAO signal with decreasing redshift, BAO are believed to be a robust standard ruler. The sound horizon scale is well determined from the CMB and the scale corresponds to $\\sim 100\\hMpc$ in present time, implying that the feature is still mostly on linear scales where evolutionary and observational effects are much simpler to predict than in the nonlinear regime. Indeed, such degradation in the contrast of the BAO due to nonlinear structure growth, redshift distortions, and possibly galaxy bias has been studied since the late 90's and is relatively well understood \\citep[][]{Meiksin99,Springel05,Angulo05,SE05,White05,Eisen05,Jeong06,Crocce06b,ESW07,Huff07,Smith07,Matarrese07,Nishimichi07,Smith08,Angulo08,Crocce08,Mat08,Taka08,Sanchez08,Jeong09,Taruya09}. Nonlinearity also induces a shift of the BAO scale in the low-redshift matter distribution relative to the sound horizon scale measured from the CMB \\citep[e.g., ][]{Smith08,Crocce08,Sanchez08}. As the demand for acoustic peak accuracy moves from the current level of a few \\% to the sub-percent level of future surveys, we are now required to understand the systematics on the BAO to much better precision. While it is evident that the shift of the BAO scale depends on the choice of the estimator, most recent studies seem to lay more weight on residual shifts using optimal estimators being at a sub-percent level at $z\\sim 0$ when accounting for nonlinear structure growth and redshift distortions \\citep{Crocce08,Sanchez08,SSEW08}, or even with halo/galaxy bias \\citep{Pad09}. Certainly more studies are necessary to confirm the results and ultimately reach a general consensus. In the previous study \\citep{SSEW08} (hereafter SSEW08), we investigated effects of nonlinear evolution on BAO using a large volume (i.e., $320\\trihGpc$) of PM simulations. We found that the shift on the acoustic scale indeed increases with decreasing redshift and is less than a percent even at $z=0.3$ in redshift space. In this work, we extend the study by using high-force-resolution simulations that are generated by a new \\Nb\\ code ABACUS by Metchnik \\& Pinto (in preparation). With the new simulations, we update the evolution of the shifts on the acoustic scale due to nonlinear structure growth and redshift distortions for various force, volume, and mass resolutions. We calculate propagators, i.e., the correlations between the linear density fields and the nonlinear density fields at low redshift, which directly manifest the nonlinear damping of the BAO. We also derive the mode-coupling contribution to the power spectrum with an attempt to qualitatively relate this to the measured values of the shifts. Recently, \\citet{PadLag09} showed that the density field after reconstruction is not the linear density field at second order. We relate the propagator and the mode-coupling term to an average and a dispersion of nonlinear density fields in the complex Fourier plane before and after reconstruction, relative to the linear density fields, and derive a signal-to-noise ratio of the standard ruler test from these two terms. The effect of galaxy bias will be presented in companion papers \\citep[Mehta et al. in preparation;][]{Xu09}. It has been demonstrated that the original density-field reconstruction scheme based on the \\Zel\\ approximation, presented by \\citet{ESSS07}, is quite efficient for removing nonlinear degradation on BAO despite its simplicity \\citep{SE07,Huff07,SSEW08,PadLag09}. Efficiency of reconstruction in terms of an increase in the signal-to-noise depends on the redshift and the shot noise of the density fields. Meanwhile, it has been shown that the reconstruction removes almost all of the nonlinear shifts of the acoustic scale even when it seemingly is not efficient in terms of the signal-to-noise ratio (SSEW08). This success, conversely, can be interpreted that a further improvement on the reconstruction scheme will be only a second order effect, at least for BAO. Nevertheless, we discuss and test possible improvements on our fiducial scheme. While there are other sophisticated reconstruction methods in the literature aimed for the recovery of velocity fields and the initial density fields \\cite[e.g.,][]{MAK06}, in this paper we limit ourselves to mild modifications to our fiducial method, mainly due to its proven success for BAO. We first test an implementation of 2LPT instead of the \\Zel\\ approximation \\citep{Zel70} and second, test iterative operations of the fiducial method in order to improve the signal-to-noise ratio. The importance of an accurate prediction of the signal-to-noise ratio for future BAO surveys is evident. The calibration of the Fisher matrix-based estimations against the \\Nb\\ results have been tried repeatedly, and the resulting discrepancy is at most 20\\% (e.g., SSEW08). Further calibration is often limited by the volume of the simulations: in order to minimize {\\it the dispersion of the dispersion} among different measurements, we need a large number of random subsamples while each subsample has an enough cosmic volume to measure the BAO scale. In this paper, we calibrate the Fisher matrix estimation based on \\citet{SE07} to a level of 1\\% by utilizing the enormous cosmic volume of $5000\\trihGpc$ from \\citet{Taka09}. This paper is organized as following. In \\S~\\ref{sec:smethod}, we describe our new cosmological \\Nb\\ simulations and the methods of $\\chi^2$ analysis to measure the acoustic scales from the simulations. In \\S~\\ref{sec:salphas}, we present the resulting shifts and errors on the measurements of the acoustic scale when accounting for the nonlinear growth and redshift distortions before and after reconstruction. In \\S~\\ref{sec:spropagators} and \\S~\\ref{sec:Pmc}, we derive propagators and mode coupling terms from the simulations and qualitatively compare these with the errors and the mean values of the shifts from the simulations. In \\S~\\ref{sec:StoN}, we relate the propagator and the mode-coupling term to the signal-to-noise ratio of the standard ruler test. In \\S~\\ref{sec:stLPT}, we implement 2LPT and iteration into our reconstruction scheme and discuss the effect. In \\S~\\ref{sec:sRyuichi}, we use the $5000\\trihGpc$ of simulations by \\citet{Taka09} to test the Fisher matrix calculations in \\citet{SE07} given the minimal sample variance. Finally, in \\S~\\ref{sec:sdisc}, we summarize the major results presented in this paper. ", "conclusions": "\\label{sec:sdisc} We summarize the results presented in this paper. 1. We have measured shifts of the acoustic scale due to nonlinear growth and redshift distortions using three sets of simulations, G576, G1024, and T256. which differ in their force, volume, and mass resolution. The measured shifts from the various simulations are in agreement within $\\sim 1.5\\sigma$ of sample variance. We numerically find $\\al(z) -1 = (0.295\\pm 0.075)\\% [D(z)/D(0)]^{1.74\\pm 0.35}$ based on G576. If we fix the power index to be 2, as expected from the perturbation theory, we find the best fit of $\\al(z) -1 = (0.300\\pm 0.015)\\% [D(z)/D(0)]^{2}$. 2. We find a strong correlation with a non-unity slope between shifts in real space and redshift space. Meanwhile the correlation with the shifts at the initial redshift is weak. Reconstruction not only removes the mean shift and reduces errors on the mean, but also tightens these correlations. After reconstruction, we recover a slope of near unity for the correlation between the shifts in real and redshift space, and restore a strong correlation between the shifts at the low and the initial redshifts. We believe that, as the reconstruction removes shifts due to the second-order, nonlinear process in structure growth, the remaining shifts are dominated by the initial conditions. 3. We find that a propagator is well described by the \\Zel\\ approximation: for $z \\lesssim 1$, we find that the discrepancy, $\\Delta \\Signl$ is less than $0.5\\hMpc$. At high redshift, \\Zel\\ approximation seems to slightly underestimate the amount of nonlinear damping. 4. We have compared our measurements of shifts from $\\chi^2$ fitting with an oscillatory feature in $\\Pmc$ and find a qualitative agreement. 5. We construct the signal-to-noise ratio of the standard ruler test based on the measured propagator and mode-coupling term assuming a Gaussian error, and find a consistency with the measured errors on shift. We point out that the propagator describes the average projection of the nonlinear density field onto the linear density field while the mode-coupling term describes any random dispersion uncorrelated to the information of the linear density field. In the case of BAO, the recovery of BAO signal is well-described by $G(k)$, therefore, the average projection on the linear field, while the mode-coupling term contributes to the noise. On the other hand, the recovery of the broadband shape of the linear power spectrum will be hard to characterize with $G(k)$ alone, in the presence of $\\Pmc$. 6. We have attempted to improve our reconstruction method by implementing 2LPT and iterative operations. We find only a few \\% improvement in $G(k)$. We will further investigate variations in the 2LPT implementation in future work. 7. We test Fisher matrix estimates of the uncertainty in the acoustic scale using $5000\\trihGpc$ of cosmological PM simulations (T256). At an expected sample variance level of 1\\%, the agreement between the Fisher matrix estimates based on \\citet{SE07} and the \\Nb\\ results is better than 10 \\%. \\\\ To conclude this paper, the acoustic peak shifts are small and can be accurately predicted, with control exceeding that required of the observational cosmic variance limit. Moreover, reconstruction removes the shifts, decreases the scatter, and improves the detailed agreement with the initial density field, as hoped. We have validated that the acoustic scale shift can be removed to better than 0.02\\% for the redshift-space matter field. We next plan to investigate the effects of galaxy bias (Mehta et al., in preparation). \\\\" }, "0910/0910.0416_arXiv.txt": { "abstract": "Spectral data are presented for comets 2006~VZ13~(LINEAR), 2006~K4~(NEAT), 2006~OF2~(Broughton), 2P/Encke, and 93P/Lovas~I, obtained with the Cerro-Tololo Inter-American Observatory 1.5-m telescope in August 2007. Comet 2006~VZ13 is a new Oort cloud comet and shows strong lines of CN (3880~\\AA{}), the Swan band sequence for C$_2$ (4740, 5160, and 5630~\\AA{}), C$_3$ (4056~\\AA{}), and other faint species. Lines are also identified in the spectra of the other comets. Flux measurements of the CN, C$_2(\\Delta v=+1,0)$, and C$_3$ lines are recorded for each comet and production rates and ratios are derived. When considering the comets as a group, there is a correlation of C$_2$ and C$_3$ production with CN, but there is no conclusive evidence that the production rate ratios depend on heliocentric distance. The continuum is also measured, and the dust production and dust-to-gas ratios are calculated. There is a general trend, for the group of comets, between the dust-to-gas ratio and heliocentric distance, but it does not depend on dynamical age or class. Comet 2006~VZ13 is determined to be in the carbon-depleted (or Tempel~1 type) class. ", "introduction": "The icy nature of comets indicate they have been preserved at cold temperatures since the early stages of solar system formation. Consequently, they are commonly considered to be among the most primitive objects in the solar system. Determining their physical and chemical properties, and how they evolve, is important to our understanding of the formation of planetary systems, both our own and in general. Several surveys have attempted to study and classify the chemical composition and evolution of comets. \\citet{A1995} conducted a photometric survey of 85 different comets over almost 20 years. Their findings indicated two major classes of comets: those that are carbon-depleted and those that are not. They found nearly all members of the carbon-depleted class are Jupiter family comets (JFCs), although not all JFCs are carbon-depleted. They also reported little variation of relative production rates with heliocentric distance or apparition; however, they noted a correlation between the dust-to-gas ratio and perihelion distance. Three major spectroscopic surveys have also been conducted: \\citet{NS1989} reported spectrophotometry of 25 comets; \\citet{C1992} derived production rates for 17 faint comets; and \\citet{FH1996} surveyed the spectra of 39 comets into infrared wavelengths. The three surveys found slightly different results. \\citet{NS1989} reported a correlation between CN and dust, and that the C$_2$/CN production ratio changed with heliocentric distance. \\citet{C1992}, however, found the gas production ratios remained constant with activity level and heliocentric distance, except for NH$_2$/CN. \\citet{FH1996} concluded that most comets have roughly the same production rate ratios to within a factor of 2 or 3, although 10~per~cent of comets could be considered outliers. More recently, \\citet{F2009} presented a spectroscopic survey of 92 comets over approximately 19 years. They report four taxanomic classes of comets: typical, Tempel~1 type, Giacobini-Zinner type, and the unusual object Yanaka (1988r). The typical comets have typical ratios of C$_2$, NH$_2$, and CN with respect to water, while Tempel-1 types have deficiencies in C$_2$ but normal NH$_2$ abundances. Giacobini-Zinner comets have low C$_2$ and NH$_2$ ratios with respect to water, while Yanaka has no detectable C$_2$ or CN emission, but normal NH$_2$ abundances. They conclude that the Halley family of comets (originating in the Saturn and Uranus region, but were scattered to the Oort cloud) shows no C$_2$ depletion, while objects originating in the Neptune region show a mixture of typical and C$_2$ depleted objects. Comets originating in the classical Kuiper belt form the C$_2$ depleted group. Comet 2006~VZ13~(LINEAR) was discovered in November 2006 \\citep{S2006}, and is a new Oort cloud comet which passed perihelion on 10 August 2007. This presented a unique opportunity to observe a pristine comet as it passed perihelion. In addition, four other comets were observed with the same instrument and under the same observing conditions: 2006~K4~(NEAT), 2006~OF2~(Broughton), 93P/Lovas~I, and 2P/Encke. These comets represent a broad range of dynamical class, age, brightness and heliocentric distance. This allows for a comparison of production rates and ratios between comets of different classes and ages. \\begin{table*} \\centering \\begin{minipage}{150mm} \\caption{Orbital elements of the comets.\\setcounter{footnote}{\\value{footnote}}\\protect\\footnotemark}\\label{tab:orbits} \\begin{tabular}{@{}lcccccccc@{}} \\hline Comet & $a$ (au) & $e$ & $i$ ($^o$) & $q$ (au) & $Q$ (au) & $T_J$ & Perihelion Date (UT) & Comet Class\\setcounter{footnote}{\\value{footnote}}\\protect\\footnotemark \\\\ \\hline 2P/Encke\t&\t2.217\t&\t0.847\t&\t11.766\t&\t0.339\t&\t4.095\t& 3.026 & 2007 Apr 19 & NEO\\\\ 2006 K4\t\t& 1805.756 &\t0.998\t&\t111.333\t&\t3.189\t&\t3608.323 &\t-0.802 & 2007 Nov 29 & EOC\\\\ 2006 OF2\t& -3367.756 &\t1.001\t&\t30.171\t&\t2.431\t&\tn/a \t& \tn/a & 2008 Sep 15 & NOC\\\\ 2006 VZ13\t& -4083.355 &\t1.000\t&\t134.793 &\t1.015 &\tn/a\t&\tn/a & 2007 Aug 10 & NOC\\\\ 93P/Lovas~1\t& 4.391\t&\t0.612\t&\t12.218\t&\t1.705\t&\t7.078\t& 2.605 & 2007 Dec 17 & JFC\\\\ \\hline \\end{tabular} \\footnotetext{$^1$ All values are from the JPL Small-Body Database Browser; $a$ is the semi-major axis, $e$ is the eccentricity, $i$ is the inclination, $q$ is the perihelion distance, $Q$ is the aphelion distance, and $T_J$ is the Jupiter Tisserand parameter (given by $T_J=a_J/a +2\\cos i\\sqrt{a(1-e^2)/a_J}$).} \\footnotetext{$^2$ NEO = near earth object; EOC = evolved Oort cloud comet; NOC = new Oort cloud comet; JFC = Jupiter family comet.} \\end{minipage} \\end{table*} \\setcounter{footnote}{0} \\begin{table*} \\centering \\begin{minipage}{105mm} \\caption{CTIO observational parameters.\\setcounter{footnote}{\\value{footnote}}\\protect\\footnotemark}\\label{tab:obs} \\begin{tabular}{@{}lccccc@{}} \\hline Comet & Date (UT) & $\\Delta$ (au) & $r_H$ (au) & $V$ (mag) & Exp. Time (s) \\\\ \\hline 2P\t\t&\t07/08/07 00:16\t&\t0.9400\t&\t1.9114\t&\t15.19\t&\t3600\\\\ &\t07/08/07 23:56\t&\t0.9550\t&\t1.9221\t&\t15.30\t&\t3650\\\\ &\t13/08/07 01:30\t&\t1.0381\t&\t1.9781\t&\t15.92\t&\t5400\\\\ &\t14/08/07 01:14\t&\t1.0551\t&\t1.9890\t&\t16.04\t&\t5400\\\\ 2006 K4\t\t&\t07/08/07 01:24\t&\t2.6270\t&\t3.3713\t&\t15.75\t&\t3600\\\\ &\t08/08/07 02:08\t&\t2.6364\t&\t3.3681\t&\t15.75\t&\t3600\\\\ &\t13/08/07 03:33\t&\t2.6869\t&\t3.3529\t&\t15.78\t&\t5400\\\\ &\t14/08/07 03:04\t&\t2.6977\t&\t3.3500\t&\t15.78\t&\t5400\\\\ &\t15/08/07 02:33\t&\t2.7084\t&\t3.3472\t&\t15.79\t&\t5400\\\\ 2006 OF2\t&\t07/08/07 02:53\t&\t3.7934\t&\t4.7860\t&\t16.33\t&\t3600\\\\ &\t08/08/07 03:19\t&\t3.7825\t&\t4.7782\t&\t16.32\t&\t3600\\\\ &\t13/08/07 06:32\t&\t3.7316\t&\t4.7382\t&\t16.26\t&\t5400\\\\ &\t14/08/07 06:17\t&\t3.7226\t&\t4.7304\t&\t16.25\t&\t5400\\\\ &\t15/08/07 05:06\t&\t3.7143\t&\t4.7229\t&\t16.24\t&\t5400\\\\ 2006 VZ13\t&\t06/08/07 23:14\t&\t1.0207\t&\t1.0175\t&\t13.26\t&\t1800\\\\ &\t07/08/07 23:14\t&\t1.0482\t&\t1.0165\t&\t13.32\t&\t1800\\\\ &\t12/08/07 23:22\t&\t1.1866\t&\t1.0159\t&\t13.58\t&\t3600\\\\ &\t13/08/07 23:19\t&\t1.2142\t&\t1.0166\t&\t13.64\t&\t3600\\\\ 93P\t\t&\t08/08/07 06:03\t&\t1.4603\t&\t2.1451\t&\t16.61\t&\t7200\\\\ &\t15/08/07 07:36\t&\t1.3617\t&\t2.1059\t&\t16.38\t&\t9600\\\\ \\hline \\end{tabular} \\footnotetext{$^1$ The date indicates the start time of the first observation of the comet; $\\Delta$ is the geocentric distance, $r_H$ is the heliocentric distance, and $V$ is the total magnitude. The exposure time is the total of all observations on the given night.} \\end{minipage} \\end{table*} In this paper, spectroscopic observations are presented of five comets, obtained in August 2007. The production rates of CN, C$_2$, and C$_3$ are calculated, and the production ratios with respect to CN are derived. In addition, dust production rates and dust-to-gas ratios are calculated. ", "conclusions": "Spectroscopic observations for five comets are reported. From these spectra, the gas production rates and production rate ratios have been calculated. There seems to be a linearity of the production rate ratios with respect to CN, agreeing with past studies \\citep{C1987,A1995}. There does not seem to be a correlation between the production rate ratios and heliocentric distance. By comparing the spectra of comet 2006~VZ13 with those presented in \\citet{F2009}, it is determined to be a carbon-depleted (Tempel~1 type) comet. The dust production rate and the dust-to-gas ratio are also calculated for each comet. The ratio stays relatively constant for each comet, except for 2P due to a variability in CN production. There seems to be an overall dependence on heliocentric distance when considering the comets as a group. There is no observed dependence of the dust-to-gas ratio on the dynamical age or class of the comets." }, "0910/0910.5719_arXiv.txt": { "abstract": "White dwarf--neutron star binaries generate detectable gravitational radiation. We construct Newtonian equilibrium models of corotational white dwarf--neutron star (WDNS) binaries in circular orbit and find that these models terminate at the Roche limit. At this point the binary will undergo either stable mass transfer (SMT) and evolve on a secular timescale, or unstable mass transfer (UMT), which results in the tidal disruption of the WD. The path a given binary will follow depends primarily on its mass ratio. We analyze the fate of known WDNS binaries and use population synthesis results to estimate the number of LISA-resolved galactic binaries that will undergo either SMT or UMT. We model the quasistationary SMT epoch by solving a set of simple ordinary differential equations and compute the corresponding gravitational waveforms. Finally, we discuss in general terms the possible fate of binaries that undergo UMT and construct approximate Newtonian equilibrium configurations of merged WDNS remnants. We use these configurations to assess plausible outcomes of our future, fully relativistic simulations of these systems. If sufficient WD debris lands on the NS, the remnant may collapse, whereby the gravitational waves from the inspiral, merger, and collapse phases will sweep from LISA through LIGO frequency bands. If the debris forms a disk about the NS, it may fragment and form planets. ", "introduction": "The inspiral and merger of compact binaries represent some of the most promising sources of gravitational waves (GWs) for detection by ground-based laser interferometers like LIGO \\cite{LIGO1,LIGO2}, VIRGO \\cite{VIRGO1,VIRGO2}, GEO \\cite{GEO}, TAMA \\cite{TAMA1,TAMA2} and AIGO \\cite{AIGO}, as well as by proposed space-based interferometers like LISA \\cite{LISA} and DECIGO \\cite{DECIGO}. Extracting physical information from gravitational radiation emitted by compact binaries requires careful modeling of these systems (see \\cite{BSReview} for a review). Most effort to date has focused on modeling black hole--black hole (BHBH) binaries \\cite{Pretorius2005a,Baker2006,Campanelli2006,Brugmann2006a,Herrmann2006,Sperhake2006,Etienne2007a,Healy,Baker2006b,Lousto2007,Gonzalez08,Scheel2008} and neutron star--neutron star (NSNS) binaries \\cite{2004PhRvD..69l4036F,2005A&A...431..297B,2005PhRvD..71h4021S,2006PhRvD..73f4027S,2008PhRvD..77b4006A,2008PhRvD..78b4012L,2008PhRvD..78h4033B}, with some recent work on black hole--neutron star (BHNS) binaries \\cite{Rantsiou08, Loffler06, Faber06, Shibata07, Shibata08,Yamamoto08, Etienne08a, Etienne08, Duez08}. In this paper we begin to explore WDNS binaries. They are plausible sources of low-frequency GWs for LISA, DECIGO and possibly also, as we shall see, high-frequency GWs for LIGO, VIRGO, GEO, TAMA and AIGO. Like NSNS binaries, WDNS binaries are known to exist. We have compiled a list of observed WDNS binaries % in Tables \\ref{table1} and \\ref{table2}. Table~\\ref{table1} lists those binaries with relatively well-determined individual component masses. By ``relatively well-determined\" we mean that in determining the masses no assumption was made about the mass of the NS. Table~\\ref{table2} lists those binaries for which the masses of the individual components have not been well-determined as yet. For these cases the NSs have been assumed to have a mass of either $1.35 M_\\odot$ or $1.40 M_\\odot$. Of all the objects presented in Tables \\ref{table1} and \\ref{table2} only B2303+46, J1141-6545, J0751+1807 and J1757-5322 have been positively identified as a pulsar with a WD companion. For all other systems there is at best strong evidence that the companion of the observed pulsar is a WD. The frequency of GWs emitted by the binary with the shortest period, J1141$-$6545, is $\\simeq 1.2\\times 10^{-4} \\rm Hz$. This frequency lies in the expected band of LISA, but well below the cutoff frequency of $\\simeq 10^{-3} \\rm Hz$ due to the double WD confusion background \\cite{Nelemans01}. It would therefore be impossible to resolve this signal. Assuming the quadrupole approximation for a WDNS binary with masses $M_{\\rm WD}$, $M_{\\rm NS}$, reduced mass $\\mu$, and total mass $M_{\\rm T}=M_{\\rm NS}+M_{\\rm WD}$ in circular orbit with angular frequency $\\Omega$, the amplitude of GWs coming from this object is $h = 4 \\mu M_{\\rm T} / (r A)$ \\cite{Shapiro}, where $r$ is the distance to the object and $A$ the binary separation. In \\cite{distanceJ11} a lower bound is given to the distance to PSR J1141$-$6545 of $r\\geq 3.7$ kpc. Using the data of Table \\ref{table1} we find that the maximum amplitude of the expected waves coming from this object is $h_{max}=1.36\\times 10^{-23}$. A gravitational wave of this amplitude is below LISA's sensitivity at $10^{-4} \\rm Hz$ ($h_s\\sim 10^{-21}$ ) and hence undetectable. In order for detectable gravitational wave signals to be emitted the orbital separation of PSR J1141$-$6545 has to decrease by a factor of about 100. The emission of gravitational radiation will cause a binary to inspiral to a close, nearly circular orbit. In this regime the binary can be described by a sequence of quasiequilibrium configurations. Once the separation of a WDNS binary becomes a few times larger than the size of the WD, the amplitude of the emitted GWs, as we will see, is large enough to allow for detection by space-based gravitational wave observatories. An important issue to address then is the number of Galactic WDNS binaries that could be detected by space-based interferometers. \\begin{table*}[t] \\caption{WDNS binaries with well-determined masses. From left to right the columns give the name of the object, the orbital period, the corresponding quadrupole GW frequency, the WD mass, the NS mass, the total mass, the mass ratio $q=M_{\\rm WD}/M_{\\rm NS}$ and the stability of mass transfer after its onset.} \\begin{tabular}{lccccccc}\\hline\\hline \\multicolumn{1}{p{2.15 cm}}{\\hspace{0.3 cm} PSR } & \\multicolumn{1}{p{2.cm}}{\\hspace{0.25 cm} $T$ (days)} & \\multicolumn{1}{p{3.5cm}}{\\hspace{0.5 cm} $f_{\\rm GW}=\\frac{2}{T}$ $(10^{-4} \\rm Hz)$ \\quad} & \\multicolumn{1}{p{2cm}}{ \\hspace{0.35 cm}$M_{\\rm WD}(M_{\\odot})$ \\qquad} & \\multicolumn{1}{p{2cm}}{\\hspace{0.3 cm} $M_{\\rm NS}(M_{\\odot})$ \\qquad } & \\multicolumn{1}{p{2cm}}{\\hspace{0.3 cm} $M_{\\rm T}(M_{\\odot})$ \\qquad } & \\multicolumn{1}{p{1.5cm}}{\\hspace{0.5 cm} $q$ \\quad} & \\multicolumn{1}{p{1.5cm}}{\\hspace{0.15 cm} Stable? \\quad} \\vspace{0.05cm} \\\\ \\hline B2303$+$46\\footnotemark[1] \t& $12.34$ & $0.0187$ & $1.3$ & $ 1.34$ & $2.64$ & $0.97$ & No \\\\ \\hline J0621$+$1002\\footnotemark[2] \t& $8.32$ & $0.0278$ & $0.67$ & $ 1.70$ & $2.37$ & $0.394$ & ? \\\\ \\hline J1141$-$6545\\footnotemark[3]$^,$\\footnotemark[4]& $0.198$ & $1.169$ & $1.02$ & $ 1.27$ & $2.29$ & $0.803$ & No \\\\ \\hline B1516$+$02B\\footnotemark[5] \t& $6.858$ & $0.0337$ & $0.13$ & $ 2.08$ & $2.21$ & $0.0625$ & Yes \\\\ \\hline J1713$+$0747\\footnotemark[1] \t& $67.8$ & $0.0034$ & $0.33$ & $ 1.60$ & $1.93$ & $0.206$ & ? \\\\ \\hline B1855$+$09\\footnotemark[1] \t & $12.3$ & $0.0188$ & $0.267$ & $ 1.58$ & $1.847$ & $0.169$ & Yes \\\\ \\hline J0437$-$4715\\footnotemark[1] \t& $5.74$ & $0.0403$ & $0.236$ & $ 1.58$ & $1.816$ & $0.149$ & Yes \\\\ \\hline J1012$+$5307\\footnotemark[1] \t& $0.605$ & $0.382$ & $0.16$ & $ 1.64$ & $1.80$ & $0.097$ & Yes \\\\ \\hline J0751$+$1807\\footnotemark[1] \t& $0.263$ & $0.88$ & $0.125$ & $ 1.26$ & $1.385$ & $0.099$ & Yes \\\\ \\hline\\hline \\end{tabular} \\begin{flushleft} \\footnotetext[1]{Stairs \\cite{WDNS7} and references therein.} % \\footnotetext[2]{Nice et al. \\cite{WDNS2}.} % \\footnotetext[3]{Bhat et al. \\cite{WDNS4}.} % \\footnotetext[4]{Bailes et al. \\cite{WDNS5}.} % \\footnotetext[5]{Freire et al. \\cite{WDNS6}.} % \\end{flushleft} \\label{table1} \\end{table*} \\begin{table*} [t] \\caption{WDNS binaries which do not have well-determined masses. From left to right the columns give the name of the object, the orbital period, the corresponding quadrupole GW frequency, the WD mass, the NS mass, the total mass, the mass ratio $q=M_{\\rm WD}/M_{\\rm NS}$ and the stability of mass transfer after its onset.} \\begin{tabular}{lccccccc}\\hline\\hline \\multicolumn{1}{p{2.15 cm}}{\\hspace{0.3 cm} PSR } & \\multicolumn{1}{p{2.cm}}{\\hspace{0.25 cm} $T$ (days)} & \\multicolumn{1}{p{3.5cm}}{\\hspace{0.5 cm} $f_{\\rm GW}=\\frac{2}{T}$ $(10^{-4} \\rm Hz)$ \\quad} & \\multicolumn{1}{p{2cm}}{ \\hspace{0.35 cm}$M_{\\rm WD}(M_{\\odot})$ \\qquad} & \\multicolumn{1}{p{2cm}}{\\hspace{0.3 cm} $M_{\\rm NS}(M_{\\odot})$ \\qquad } & \\multicolumn{1}{p{2cm}}{\\hspace{0.3 cm} $M_{\\rm T}(M_{\\odot})$ \\qquad } & \\multicolumn{1}{p{1.5cm}}{\\hspace{0.5 cm} $q$ \\quad} & \\multicolumn{1}{p{1.5cm}}{\\hspace{0.15 cm} Stable? \\quad} \\vspace{0.05cm} \\\\ \\hline J1435$-$60\\footnotemark[1] \t& $1.355$ & $0.1708$ & $1.10$ & $ 1.40$ & $2.50$ & $0.785$ & No \\\\ \\hline J1157$-$5114\\footnotemark[2] \t& $3.507$ & $0.066$ & $1.14$ & $ 1.35$ & $2.49$ & $0.844$ & No \\\\ \\hline J1453$-$58\\footnotemark[1] \t& $12.42$ & $0.0186$ & $1.07$ & $ 1.40$ & $2.47$ & $0.764$ & No \\\\ \\hline J1022$+$1001\\footnotemark[1] & $7.805$ & $0.0296$ & $0.872$ & $ 1.40$ & $2.272$ & $0.623$ & No \\\\ \\hline B0655$+$64\\footnotemark[1] \t& $1.029$ & $0.2948$ & $0.814$ & $ 1.40$ & $2.214$ & $0.581$ & No \\\\ \\hline J2145$-$0750\\footnotemark[1] \t& $6.839$ & $0.0338$ & $0.515$ & $1.40$ & $1.915$ & $0.368$ & ? \\\\ \\hline J1757$-$5322\\footnotemark[2] \t& $0.453$ & $0.511$ & $0.55$ & $1.35$ & $1.90$ & $0.407$ & ? \\\\ \\hline J1603$-$7202\\footnotemark[1] \t& $6.309$ & $0.0367$ & $0.346$ & $ 1.40$ & $1.746$ & $0.247$ & ? \\\\ \\hline J1810$-$2005\\footnotemark[1] & $15.01$ & $0.0154$ & $0.34$ & $ 1.40$ & $1.74$ & $0.243$ & ? \\\\ \\hline J1904$+$04\\footnotemark[1] \t& $15.75$ & $0.0147$ & $0.27$ & $ 1.40$ & $1.67$ & $0.193$ & ? \\\\ \\hline J1232$-$6501\\footnotemark[1] & $3.507$ & $0.066$ & $0.175$ & $ 1.40$ & $1.575$ & $0.125$ & Yes \\\\ \\hline\\hline \\end{tabular} \\begin{flushleft} \\footnotetext[1]{Tauris et al. \\cite{WDNS1} and references therein.} % \\footnotetext[2]{Edwards and Bailes \\cite{WDNS3}.} % \\end{flushleft} \\label{table2} \\end{table*} Population synthesis calculations by Nelemans et al.\\ \\cite{Nelemans01} show that there are about $2.2\\times 10^{6}$ WDNS binaries in our Galaxy, and that they have a merger rate of $1.4\\times 10^{-4} \\rm yr^{-1}$. Furthermore, Nelemans et al.\\ find that after a year of integration, LISA should be able to detect $128$ WDNS binaries and, after considering the contribution of the double WD background GW noise, resolve $38$ of these. On the other hand, calculations by Cooray \\cite{Cooray2004}, give much more conservative numbers of resolved WDNS binaries. In particular, Cooray finds that the number of LISA-resolved WDNS binaries ranges between $1$--$10$, using a WDNS merger rate between $10^{-6} \\rm yr^{-1}$--$10^{-5} \\rm yr^{-1}$. Cooray's upper limit was based on merger rates calculated by Kim et al.\\ \\cite{Kim2004}. In this work we focus on WDNS binaries in close binary separations, and examine the termination point of quasiequilibrium sequences describing such binaries. Several different astrophysical scenarios can result in such a termination point for binaries in general, including {\\em direct contact} of the binary components, {\\em Roche lobe overflow} by one of the two companions, or the binary reaching an {\\em innermost stable circular orbit} (ISCO). As we will find, quasiequilibrium sequences of WDNS terminate when the WD fills its Roche lobe, resulting in mass transfer from the WD onto the NS across the inner Lagrange point. This mass transfer can either be stable (SMT) or unstable (UMT). We also refer to the latter scenario as the tidal disruption of the WD by the NS. To determine which of these two outcomes is likely for a given system, we will follow the approach of Verbunt and Rappaport \\cite{Verbunt88} and Faber et al. \\cite{Faber}. As indicated in our Tables~\\ref{table1} and~\\ref{table2}, we shall find that among the observed WDNS binaries, some will undergo SMT, while others will undergo tidal disruption (i.e.~UMT). Note that mass transfer stability has also been studied in \\cite{Hut,Marsh}. In the case of SMT, the orbital evolution of the binary occurs on a secular timescale determined by the emission of gravitational radiation. Therefore, the quasistationary conservative treatment of Clark and Eardley \\cite{Clark77} and Faber et al. \\cite{Faber} is adequate to follow the evolution during this secular phase. Note that a quasistationary treatment to follow the orbital evolution of binary systems has been employed by several authors in the past. For example, Rappaport et al.~\\cite{Rappaport82} studied compact binaries where the mass of the secondary can be up to $\\sim 1M_\\odot$ and modeled as a $n=3/2$ polytrope. Fryer et al.~\\cite{Fryer99} employed a non-conservative quasistationary approach to study the evolution of white dwarf--black hole binaries, while Marsh et al.~\\cite{Marsh} studied the evolution of double WD binary systems. Both the tidal disruption and SMT phase of double WD systems have also been studied in \\cite{Benz1990, Podsiadlowski92, RasioShapiro95, Segretain1997, Guerrero2004, Yoon2007,Dan08} via SPH simulations and in \\cite{Dsouza,Motl} via grid-based hydrodynamic calculations, all in Newtonian gravitation. Finally, Newtonian SPH simulations of encounters of WDs with intermediate mass BHs ($M_{\\rm BH}\\sim 10^3 M_\\odot$) have been performed in \\cite{RosswogWDBH1,RosswogWDBH2}. In the case of tidal disruption, on the other hand, the system will evolve on a hydrodynamical (orbital) timescale. In this scenario the NS may plunge into the WD and spiral toward the center of the star liberating its gravitational potential energy as heat in the WD material. Alternatively, the NS may be the receptacle of massive debris from the disrupted WD. Depending on the details of the equation of state, a cold degenerate gas can support a maximum NS rest-mass between $1.89 M_\\odot-2.67 M_\\odot$ (corresponding to a gravitational mass between $1.65M_\\odot-2.20M_\\odot$) against catastrophic collapse if it is not rotating (the OV limit), about $20\\%$ more mass if it is rotating uniformly (a ``supramassive NS'', e.g.~\\cite{CooST94}), and at least $50\\%$ more mass if it rotates differentially (a ``hypermassive NS'') \\cite{BauSS00,2004ApJ...610..941M}. The fate of the merged WDNS then depends on the initial masses of the progenitor stars, the degree of mass and angular momentum loss during the WD disruption and binary merger phases, the angular momentum profile of the NS remnant and the extent to which the remnant gas is heated by shocks as it pours onto the NS and forms an extended, massive mantle. These are issues that require a hydrodynamic simulation to resolve. Moreover, ascertaining whether or not the neutron star ultimately undergoes a catastrophic collapse to a black hole (either prompt or delayed) requires that such a simulation be performed in full general relativity. We plan to explore some of these alternative hydrodynamical scenarios in detail in the future, aided by simulations that employ our adaptive mesh refinement (AMR) relativistic hydrodynamics code \\cite{Etienne08}. In this paper we survey the problem in qualitative terms. In Section~\\ref{Sec:Eq} we identify some of physical parameters that are likely to play a key role in the evolution of a WDNS binary system. We then model the secular inspiral epoch of the binary by constructing Newtonian equilibrium models of corotational WDNS binaries in close circular orbit, up to the Roche limit. In Section~\\ref{MTSvsTD} we follow the stability analysis of Verbunt and Rappaport \\cite{Verbunt88} in order to determine the late-evolution of these binaries (SMT vs.~tidal disruption). We treat the quasistationary SMT epoch by applying the approach of Clark and Eardley \\cite{Clark77} and Faber et al.~\\cite{Faber} and compute the corresponding gravitational waveforms. We discuss in general terms the possible outcomes of binaries that undergo UMT and construct approximate equilibrium configurations of merged WDNS remnants. We use these models to make some predictions regarding our future fully relativistic simulations of these systems. Finally, we conclude in Section \\ref{summary}, where we summarize the main findings of this work. ", "conclusions": "\\label{summary} In this paper we considered WDNS binaries and ``set the stage\" for fully relativistic hydrodynamic simulations of the WDNS merger. Like NSNS binaries, but unlike BHNS and BHBH binaries, WDNS binaries are known to exist and are abundant. We have compiled a list of observed WDNS binaries and their properties in Tables \\ref{table1} and \\ref{table2}. The emission of gravitational radiation will cause the orbital separation of a WDNS binary to shrink. We modeled corotational WDNS binaries in circular orbit and found that these models terminate at the Roche limit. At this point the binary can undergo either SMT and evolve on a secular timescale or UMT and evolve on a hydrodynamical timescale. Following the stability analysis of Verbunt and Rappaport \\cite{Verbunt}, we showed that the subsequent fate of the binary is determined by a critical mass ratio. Using the results of this stability analysis we predicted the possible fates of known WDNS binaries and we indicated this in our Tables~\\ref{table1} and~\\ref{table2}. Furthermore, based on population synthesis results by Nelemans et al. \\cite{Nelemans01}, we gave an estimate of the number of LISA-resolved galactic binaries per year that will undergo SMT and the number of those that will undergo tidal disruption. We found that approximately up to $16$ WDNS binaries will undergo UMT (tidal disruption), and up to $20$ SMT. In the case of SMT, the timescale of the mass transfer, and hence the timescale of the binary evolution, is set by the emission of gravitational radiation. We treated this quasistationary SMT epoch by applying the approach of Clark and Eardley \\cite{Clark77} and Faber et al. \\cite{Faber}, also adopted in \\cite{Rappaport82, Fryer99,Marsh}, and we estimated the corresponding gravitational waveforms. We also constructed approximate equilibrium configurations of rigidly rotating WDNS merged remnants in order to explore possible outcomes of WDNS mergers. We found that unless some process removes angular momentum from the system or leads to the formation of a massive disk, these massive equilibrium configurations cannot collapse directly to form a Kerr black hole since their angular momentum ($J/M^2\\sim 20>1$) exceeds the Kerr limit. Furthermore, in most of our case studies we found that the merged remnants have a ratio of spin kinetic to gravitational potential energy greater than $0.27$, which implies that they are dynamically unstable to bar formation. However, we emphasize that these preliminary results are at best approximate and must be confirmed and/or revised by detailed numerical simulations. The fate of a merged WDNS binary depends on the initial mass of the cold progenitor stars, the degree of mass and angular momentum loss during the WD disruption and binary merger phases, the angular momentum profile of the WDNS remnant and the extent to which the remnant gas is heated by shocks as it pours onto the NS and forms an extended, massive mantle. These issues all require a hydrodynamic simulation to resolve. Moreover, ascertaining whether or not the NS ultimately undergoes a catastrophic collapse to a BH (either prompt or delayed) requires that such a simulation be performed in full general relativity. In fact, even the final fate of the NS in the alternative scenario in which there is a long epoch of SMT may also lead to catastrophic collapse, if the final NS mass exceeds the TOV mass limit. This scenario too will require a general relativistic hydrodynamic simulation to track. We plan to explore some of these alternative scenarios in detail in the future, aided by simulations that employ our AMR relativistic hydrodynamics code." }, "0910/0910.5233_arXiv.txt": { "abstract": "We calculate the distribution of neutral Hydrogen within 750 proper $\\chikps$ of a quasar, $L_{\\rm bol} = 1.62 \\times 10^{13} {\\rm L}_{\\sun} \\approx L_{\\rm Edd}$, powered by accretion onto a super massive black hole, $M_{\\rm BH} = 4.47 \\times 10^8 {\\rm M}_{\\sun}$, at z = 3. Our numerical model includes cosmological initial conditions, gas dynamics, star formation, supernovae feedback, and the self consistent growth and thermal feedback of black holes calculated using \\Gadget as well as a detailed post-processing ray tracing treatment of the non-uniform ionizing radiation field calculated using \\Sphray. Our radiative transfer scheme naturally accounts for the self shielding of optically thick systems near the luminous central source. We show that the correct treatment of self shielding introduces a flattening feature into the neutral column density distribution around Log $N_{\\rm HI}$ = 20 and that regions with the lowest neutral fractions are not necessarily those with the highest density gas. For comparison with our numerical work, we solve a Ricatti equation which determines the equilibrium Hydrogen ionization fractions in the presence of a radiation field that falls off as one over $r$ squared with regions above a given gas density threshold completely shielded from ionizing radiation. We demonstrate that these simple semi analytic models cannot reproduce the neutral Hydrogen field calculated using {\\small SPHRAY}. We conclude by comparing our models of this single proximity zone to observations by Hennawi and Prochaska of the absorption spectra of background quasars which are coincident on the sky with foreground quasars in their Quasars Probing Quasars (QPQ) series of papers. Compared to the QPQ sample, we find a factor of 3 fewer optically thick (Log $N_{\\rm HI} \\ge 17.2$) systems around our quasar, however the dark matter halo that hosts our simulated quasar, $M_{\\rm Halo} = 5.25 \\times 10^{12} {\\rm M}_{\\sun}$, is less massive than the typical QPQ host halo by a factor of four. Allowing for a linear scaling between halo mass, baryonic overdensity and number of absorbers, we estimate the typical host halo mass in the QPQ sample as $1.92 \\times 10^{13} {\\rm M}_{\\sun}$. ", "introduction": "After the reionization of the Universe by the first luminous objects, the majority of neutral Hydrogen resides in gravitationally collapsed objects, specifically Damped Lyman-$\\alpha$ systems (DLAs). These systems are observed via absorption lines in the spectra of distant quasars and are historically identified as those absorbers with HI column densities $N_{\\rm HI} > 2 \\times 10^{20} {\\rm cm^{-2}}$ with lower column density systems $10^{17.2} < N_{\\rm HI} < 2 \\times 10^{20} {\\rm cm^{-2}}$ being labeled Lyman Limit Systems \\cite[LLSs, see][for a review]{2005ARA&A..43..861W}. In contrast, the systems that give rise to the Lyman-$\\alpha$ forest have $N_{\\rm HI} < 10^{17.2} {\\rm cm^{-2}}$ \\cite[see][for reviews]{1998ARA&A..36..267R, 2003AIPC..666..157W}. This historical column density threshold ($N_{\\rm HI} = 10^{17.2} {\\rm cm^{-2}}$) serves to divide systems into those that are predominantly neutral (the DLAs and LLSs) and those that reside in the mostly ionized intergalactic medium (IGM). The DLAs and LLSs remain mostly neutral in the presence of an ionizing ultraviolet (UV) background and local point sources through self shielding. Studying them in absorption opens a window into the post reionization population of cold, dense, neutral gas and provides a survey technique with a bias complimentary to that of emission surveys. To study these systems numerically requires treating the UV ionizing background at some level. Many cosmological simulations have followed in the steps of \\cite{HM96} and \\cite{1996ApJS..105...19K} by considering a spatially uniform, time variable, background with a spectral shape characteristic of quasar and stellar radiation that has been reprocessed by the IGM. The background plays a role in determining the cooling function of cosmological gas and so its inclusion is also necessary at some level for a realistic description of galaxy and star formation. A uniform background is a good first approximation if one is interested in radiative cooling, however it ignores completely the self shielding that defines the DLAs and LLSs. The most straight forward way to account for this shielding is to reduce the UV field to zero in regions above a given gas density threshold as was done in \\cite{1998ApJ...495..647H}. A more detailed treatment based on the solution for plane parallel radiation incident on a constant density slab is described in \\cite{KWHM96} and used in the work of \\cite{1997ApJ...484...31G, 2001ApJ...559..131G, 1997ApJ...486...42G}. These corrections are not based on transferring radiation through the simulation volume, but rather on applying the plane parallel solution on a pixel by pixel basis to HI column density maps in post processing. Other approaches to self shielding include that of \\cite{2003ApJ...598..741C} who include attenuation of the UV background on a cell by cell basis akin to \\cite{KWHM96}. \\cite{2006ApJ...645...55R}, who include a treatment of the transfer of ionizing radiation in post processing using the Fully Threaded Transport Engine ({\\small FTTE}) described in \\cite{2005MNRAS.362.1413R}. And \\cite{2007ApJ...660..945N}, who use a multi phase gas model to treat star formation and assume the cold dense phase to be fully neutral. Recently, a series of nested galaxy formation simulations at different resolutions was used by \\cite{2008MNRAS.390.1349P} to study DLAs over a broad range of mass scales. They include a simplified radiative transfer scheme to account for self shielding which lies somewhere between the full radiative transfer modeling and the pixel by pixel corrections. The works mentioned above were aimed at studying systems where the radiation field is not dominated by a local point source. \\cite{2005ApJ...620L..91M} applied simple analytic arguments based on the conservation of surface brightness to argue that {\\it on average} the effect from local sources is negligible compared to that of the background for systems with optical depths below that of Lyman Limit Systems. On the other hand, \\cite{2006ApJ...643...59S} use analytic arguments to show the local radiation field is likely to be important for denser systems that tend to cluster around the large scale overdensities that host these sources. In this work, we examine the balance between the UV background, local sources, and self shielding by combining, for the first time, a hydrodynamic simulation that tracks the formation and accretion history of black holes with a detailed ray tracing treatment of the non-uniform UV radiation field. Shielding is most important in the presence of dense gas and strong UV fields, both of which occur near active galactic nuclei (AGN) and quasars. In fact, the transition between the background UV field and the local AGN/quasar UV field serves to define the proximity region of these objects. With the availability of large quasar catalogues and high resolution spectroscopy, it has become possible to study absorbers proximate to a foreground quasar in the spectrum of a coincident background quasar. Work such as this has been carried out in a series of papers entitled ``Quasars Probing Quasars'', the first of which is authored by \\cite{QpQI} (HP06 in the rest of this paper). We compare our numerical models to observations from this body of work and to theoretical calculations of the HI column density by \\cite{2002ApJ...568L..71Z}. The format of this paper is as follows. In \\S2 we describe the \\Gadget simulation that determines our temperature and density fields, in \\S3 we review the ray tracing code \\Sphray used to calculate the transfer of ionizing radiation through this density field, in \\S4 we describe how we model our sources of ionizing radiation, in \\S5 we describe our semi analytic model, in \\S6 we describe our ray tracing results and compare them to the semi analytic model, in \\S7 we compare our theoretical results with those of \\cite{QpQI}, and \\cite{2002ApJ...568L..71Z} and in \\S8 we conclude and discuss further work. ", "conclusions": "In this paper, we have presented a detailed model for the neutral hydrogen gas in the proximity zone of a quasar powered by accretion onto a super massive black hole. To do this, we combined a cosmological hydrodynamic \\Gadget simulation of the formation and growth of black holes with a detailed radiative transfer post processing code {\\small SPHRAY}. We focused our attention on self shielding of the neutral gas and, by construction, the contribution from a local central source. Much of the previous work on Lyman Limit and Damped Lyman $\\alpha$ systems ignored the contribution from point sources assuming that the background field dominates. This is true, on average, for absorbers less dense then Lyman Limit systems but is not true for the more dense absorbers which cluster around luminous point sources and those underdense systems that just happen to be near sources. In addition it is crucial that these sorts of models be constructed to interpret quasar pair observations such as those in HP06. We find that the results of our ray tracing algorithm \\Sphray cannot be reproduced with simple semi analytic models in which gas above a given density threshold is completely shielded from ionizing radiation. This is because the photoionization rate at any given point is determined by the partial shielding provided by the local density field and to a lesser degree the partial shadowing seen in Figure \\ref{GHIhealpix}. We have shown in Figure \\ref{NHIpdf} that our \\Gadget + \\Sphray models naturally reproduce the observed shape of the $N_{\\rm HI}$ column density distribution. We have also shown that the flattening feature around Log $N_{\\rm HI} = 20$ attributed to self shielding in the analytic work of \\cite{2002ApJ...568L..71Z} is not due entirely to self shielding, but is largely enhanced by it. When compared to the observations of HP06, our ray tracing and semi analytic models fall short of capturing the number of dense absorbers proximate to luminous quasars. We have described improvements to our model that would bring our results into agreement with these observations including sampling more massive host halos, correcting for the under resolved clumping of gas, allowing obscuration / periodic emission of the quasar radiation and using a larger simulation volume to better model uncertain radial separations. These issues are not intractable and can be dealt with using a larger low resolution simulation to sample more massive dark matter host halos coupled with a high resolution resimulation of a more representative proximity zone. Finally we have shown that the total amount of neutral HI within the 750 ${h^{-1}}$ kpc region around our central quasar can change by up to an order of magnitude depending on the shielding prescription used. This means that HI surveys that do not resolve individual galaxy features but whose goal is to measure the integrated signal from a proximity zone will need to take these effects into account. This is on scales much smaller then the intensity mapping proposed by \\cite{2009arXiv0902.3091P} to measure Baryon Acoustic Oscillations, but would be relevant for connecting HI or 21 cm galaxy surveys to the underlying density field." }, "0910/0910.4623_arXiv.txt": { "abstract": "We present a semi-analytical method to investigate the systematic effects and statistical uncertainties of the calculated angular power spectrum when incomplete spherical maps are used. The computed power spectrum suffers in particular a loss of angular frequency resolution, which can be written as $\\delta \\ell \\sim \\pi/\\gamma_{max}$, where $\\gamma_{max}$ is the effective maximum extent of the partial spherical maps. We propose a correction algorithm to reduce systematic effects on the estimated $C_\\ell$, as obtained from the partial map projection on the spherical harmonic $\\ylm{\\ell}{m}$ basis. We have derived near optimal bands and weighting functions in $\\ell$-space for power spectrum calculation using small maps, and a correction algorithm for partially masked spherical maps that contain information on the angular correlations on all scales. ", "introduction": " ", "conclusions": "" }, "0910/0910.3235_arXiv.txt": { "abstract": "We study both the background evolution and cosmological perturbations of anisotropic inflationary models supported by coupled scalar and vector fields. The models we study preserve the $U(1)$ gauge symmetry associated with the vector field, and therefore do not possess instabilities associated with longitudinal modes (which instead plague some recently proposed models of vector inflation and curvaton). We first intoduce a model in which the background anisotropy slowly decreases during inflation; we then confirm the stability of the background solution by studying the quadratic action for all the perturbations of the model. We then compute the spectrum of the $h_{\\times}$ gravitational wave polarization. The spectrum we find breaks statistical isotropy at the largest scales and reduces to the standard nearly scale invariant form at small scales. We finally discuss the possible relevance of our results to the large scale CMB anomalies. ", "introduction": "The inflationary paradigm has become a widely accepted description of the early universe, which has been successful in solving the classical problems of FRW cosmology~\\cite{Guth,Linde}. Moreover, inflation provides a mechanism for generation of nearly scale invariant spectrum of perturbations that can seed the structure we observe today~\\cite{mfb}. Inflationary era is characterized by a quasi de-Sitter expansion with a nearly constant Hubble rate. During inflation, quantum fluctuations are generated and amplified to scales above the Hubble scale (which specifies the horizon of casual processes) where they remain frozen until they re-enter the horizon after inflation has ended. A key assumption about the quantum fluctuations during the quasi de-Sitter stage is that they are in an adiabatic vacuum stage in the deep sub-horizon (early time) regime, which results in a nearly scale invariant spectrum. Results of the WMAP experiment~\\cite{WMAP} are in overall agreement with the predictions of inflation. However, certain features of the full sky CMB maps seem to be anomalous in the standard picture. These anomalies include the low power in the quadrupole moment~\\cite{lowl}, the alignment of the lowest multipole moments (known as the 'axis-of-evil')~\\cite{axis}, and an asymmetry in power between the northern and southern ecliptic hemispheres~\\cite{asym}. The statistical significance of these anomalies has been extensively discussed in the literature and despite numerous efforts, an explanation due to a systematic effect or a foreground signal affecting the analysis is not forthcoming. These anomalies suggest a violation of statistical isotropy of cosmological fluctuations, which is at odds with standard mechanisms of generation of perturbations in inflationary models. Therefore, the possibility of relating the anomalies with modifications in the standard inflationary picture has been considered by numerous authors recently. In the standard picture, $a_{lm}$ coefficients of CMB temperature anisotropies satisfy $< a_{lm}\\, a_{lm}^* > = C_l\\, \\delta_{ll'}\\, \\delta_{mm'}$ (so, the temperature fluctuations are said to be statistically isotropic). It has been shown in~\\cite{gcp,gcp2} that an anisotropic expansion that took place at the onset of standard single field inflation, leads to a nonvanishing correlation of the $a_{lm}$ coefficients for different $l$-modes. This violation of statistical isotropy might then be related to the alignment of the lowest multipoles observed in the CMB data. Therefore, several authors considered inflationary models with an anisotropic stage. As the simplest possible modification, anisotropic initial conditions at the onset of single field inflation have been considered in references~\\cite{gcp,gcp2,uzan1,uzan2}. Isotropization of the universe takes place due to the fast expansion supported by the inflaton field. Fluctuations at scales comparable to the horizon scale during the anisotropic stage of inflation are sensitive to the background evolution and thus provide breaking of statistical isotropy at those scales. If inflation had a limited duration, this could leave an imprint on the largest observable scales. On the other hand, fluctuations that leave the horizon after isotropization re-enter earlier than modes that left the horizon during the anisotropic stage, producing the standard nearly scale invariant spectrum at small scales~\\footnote{It was also shown in~\\cite{gkp} that the anisotropic stage could lead to a detectable gravitational wave signal.}. Since in such models isotropy is reached very soon (within one Hubble time) the breaking of statistical isotropy can be visible if the total duration of inflation is tuned to only the minimal duration to solve the standard problems of FRW cosmology. The fine-tuning can be relaxed if the anisotropic stage is prolonged, due to the existence of other sources, for example vector fields. Recently, anisotropic expansion driven by vector fields have been studied by many authors. The anisotropic expansion is achieved when the vector field acquires a nonzero vacuum expectation value (VEV) along a spatial direction. There is a range of models which differ by the way the VEV is obtained. The first anisotropic inflationary model was considered long ago in reference~\\cite{ford}, where the $U(1)$ symmetry of the vector field action is broken by a potential term, which needs to have a negative curvature to support an anisotropic expansion. It has been shown in reference~\\cite{hcp1} that this model is plagued by a ghost instability, and its quantum theory is inconsistent~\\footnote{One may hope that this model has a well behaved UV completion. However, inflationary predictions are sensitive to this high energy regime, where the completion is needed.}. In another model, the ACW model~\\cite{acw}, the VEV is obtained by a lagrange multiplier field which fixes the norm of the vector field and breaks the $U(1)$ gauge symmetry. The stability analysis of the ACW model by considering the most general perturbations of the background has been performed in references~\\cite{dgw} and~\\cite{hcp1,hcp2}. In the latter studies, it has been shown that the linearized perturbations diverge close to horizon crossing, indicating the instability of the background solution. Another model~\\cite{soda1} considered a nonminimal coupling of the vector field to the scalar curvature which breaks the conformal invariance of the vector field action. The coupling with curvature allows for a slow-roll phase during the prolonged anisotropic expansion, and the universe isotropizes due to the existence of a massive scalar field, which is responsible for the overall isotropic expansion of the universe. Models of inflation with nonmininal coupling to curvature have also been considered for standard isotropic inflation~\\cite{mukhvect}. Isotropy of space is realized if $3N$ mutually orthogonal vector fields have equal VEVs. Instead, for randomly oriented $N$ vector fields, one expects a statistically isotropic background with order $1/\\sqrt{N}$ anisotropy. Perturbative calculations based around such a background configuration have been considered in references~\\cite{Golovnev:2008hv, Golovnev:2009ks, Golovnev:2009rm} and in~\\cite{hcp3}. The latter study is the only complete study linearized study of perturbations, taking into account all the physical degrees of freedom, and coupling between vector field and metric perturbations. It has been shown in~\\cite{hcp3} that the equations of motion for the linearized perturbations become singular close to horizon crossing, leading to instabilities. In all of these models (including~\\cite{acw,soda1,mukhvect}), instabilities are related to the existence of the longitudinal polarization of the vector field (which would otherwise be absent when the $U(1)$ symmetry is restored), which becomes a ghost close to horizon crossing (In the vector inflation model~\\cite{mukhvect}, ghosts also appear in the deep UV regime). More recently, reference~\\cite{Watanabe:2009ct} introduced an anisotropic inflation model driven simultaneously by a vector field and a massive scalar field. The vector field is massless, but it is coupled to the scalar field through its kinetic term. Such type of coupling preserves the $U(1)$ gauge invariance of the vector field, so that the dangerous longitudinal vector polarization is absent; moreover the conformal invariance of the vector field is broken. When the vector field has a nonvanishing VEV along a spatial direction, the model possess an anisotropically expanding attractor solution. The anisotropy in the attractor is initially small, but it increases towards the end of inflation; therefore this is a counter example to the cosmic no hair conjecture (See~\\cite{Kaloper:1991rw} for a different example) .The breaking of conformal invariance due to the scalar-vector coupling can also be used to generate magnetic fields from inflation, as discussed in reference~\\cite{Turner:1987bw} and more recently in~\\cite{Gordon:2000hv, Demozzi:2009fu} and~\\cite{Watanabe:2009ct}. These models are expected to be free from ghost instabilities as long as the coupling to the scalar field remains positive. However to our knowledge, there has been no complete study of the stability of these models, which take into account all the physical degrees of freedom of the system (including gravity). A complete study of stability is therefore necessary, given the problems identified with other vector field models~\\cite{hcp1,hcp2,hcp3}. In this work, we will present a complete study of cosmological perturbations of models with scalar-vector coupling that lead to an anisotropically expanding Bianchi-I background solution. We will also consider generalizations of the original model introduced in~\\cite{Watanabe:2009ct}, to include additional scalar fields, so that isotropization can be achieved before the end of inflation. This is required in order to obtain a scale invariant spectrum of perturbations at small scales, but the large scale spectrum is modified, which in turn can be related to the CMB anomalies. Our study has two main steps: one is to show that the model is free of ghost instabilities, and the second is to study the resulting phenomenology. We perform the first step explicitly in this paper. We develop the necessary tools to study the phenomenology of the model and as a simple exercise, we study only the spectrum of $h_{\\times}$ gravitational wave polarization in this paper. The study of the full phenomenology (including the spectrum of the curvature perturbation) will be communicated elsewhere~\\cite{me}. We will follow the formalism developed in~\\cite{gcp2,hcp2} to decompose and classify the perturbations. The generic background metric we study is the Bianchi-I metric with a residual two dimensional isotropy, given by \\begin{equation} ds^2 = -dt^2 + a(t)^2\\, dx^2 + b(t)^2\\, \\left( dy^2 + dz^2 \\right) \\nonumber \\end{equation} We exploit the two dimensional isotropy of the $y-z$ plane to decompose the perturbations into two decoupled classes: $2d$ scalar modes and $2d$ vector modes (As discussed in~\\cite{gcp2}, there is no two dimensional transverse-traceless mode). The linearized Einstein equations and the quadratic action for the two types of modes are decoupled and therefore we study them separately. We can deduce the number of degrees of freedom coming from each of the fields (gravity+vector field+scalar field) by a simple counting, which does not depend on the decomposition chosen to classify them. The metric has $10$ perturbations ($\\delta g_{\\mu\\nu}$) to start with, 4 of which can be removed by coordinate transformations. Of the remaining 6 modes, 4 are nondynamical, which can be best understood from the ADM formalism~\\cite{adm}. In the ADM formalism, the gravitational action is decomposed into the dynamical part containing the spatial metric $h_{ij}$ and a part containing the lapse (N) and shift functions ($N_i$). The lapse and shift functions have no kinetic terms in the action, and they can be integrated out by solving their equations of motion~\\footnote{The equations of motion derived from extremezing the gravitational action with respect to lapse and shift functions result in the momentum and hamiltonian constraints.}. This leaves only $h_{ij}$ as dynamical modes, which have only two degrees of freedom. In summary, the metric perturbations have only 2 dynamical degrees of freedom (which are the gravitational wave polarizations in the standard case) and 4 nondyncamical modes. For the case of the vector field, out of 4 perturbations to start with ($\\delta A_{\\mu}$), one perturbation can be removed by the $U(1)$ gauge transformation. Out of the remaining 3 perturbations 2 of them are dynamical and one mode is nondyamical ($\\delta A_0$). The scalar field perturbation introduces a single dynamical degree of freedom. Therefore, in total, the perturbations comprise of 5 dynamical modes, 2 of which are $2d$ vector and 3 are $2d$ scalar modes. When additional scalar fields are considered, each field introduces an extra scalar dynamical degree of freedom. We will insert the perturbative expansion of the metric, the vector field and the scalar field(s) into the starting action and expand it at the quadratic order. We will show that the actions for the $2d$ vector and $2d$ scalar modes are decoupled, so we study them separately. We will determine the linear combinations of perturbations that canonically normalize the action, generalizing the computation of the standard isotropic case~\\cite{musa}. The study of the quadratic action is crucial for both showing that the model is consistent (free of ghosts) and also for determining the initial conditions for the perturbations. For modes that are inside the horizon, canonical combinations of perturbations can be quantized, and initial conditions can be set by the canonical commutation relations and by the requirement that the adiabatic vacuum state has minimal energy. We then proceed by numerically integrating the evolution equations for the canonical fields, starting from adiabatic initial conditions deep inside the horizon, until the end of inflation, which gives us the primordial power spectra. We do so for the $2d$ vector modes in this paper (which will be related to the power spectrum of the $h_{\\times}$ polarization of gravitational waves) and perform the study of the spectrum of $2d$ scalar modes in a separate publication. The paper is organized as follows: In section~\\ref{sec:back}, we summarize the anisotropic inflationary background solution obtained in reference~\\cite{Watanabe:2009ct} and generalize the model to possess extra scalar fields. This generalization leads to isotropization before the end of inflation. In section~\\ref{sec:perturbations}, we discuss the perturbations around the background configuration. Specifically, in subsection~\\ref{sub:decomposition}, we review and discuss the decomposition of perturbations around the Bianchi-I background solution, following the formulation of reference~\\cite{gcp2}. In subsection~\\ref{sub:generic}, we discuss the generic properties of coupled vector-scalar models and develop tools that we will use in the following sections in order to find adiabatic initial conditions for such systems. In subsection~\\ref{sub:2dV} we discuss $2d$ vector perturbations around the background configuration. We compute the action for $2d$ vector modes and find the combinations of perturbations that canonically normalize the action. In section~\\ref{sec:power}, we compute the power spectra of the $2d$ vector modes numerically. This study results in the spectrum of $h_{\\times}$ gravitational wave mode. We show that the spectrum has angular dependence (so breaks statistical isotropy) at large scales and reduces to the standard nearly scale invariant form at small scales. We also provide a fit to the numerically obtained spectrum. Finally in section~\\ref{sec:conclusions} we provide a general discussion about the results we have obtained and their possible relation to observations. We also provide two extensive appendices: In appendix~\\ref{sub:app1}, we derive the relation between the $2d$ decomposed modes and the standard longitudinal mode, which has been used to determine the power spectra. In appendix~\\ref{sub:app2}, we perform the study of the $2d$ scalar modes. More precisely, we compute the quadratic action and determine the modes that canonically normalizes the action. The spectrum of scalar perturbations and the consequent phenomenological predictions (specifically, the $$ correlation) will be presented elsewhere. ", "conclusions": "\\label{sec:conclusions} In this work we have considered anisotropic inflationary models with coupled vector and scalar fields. The coupling of the scalar field to the vector field is chosen to preserve the $U(1)$ gauge symmetry of the Lagrangian, therefore, the longitudinal polarization of the vector field, which has been shown to cause instabilities~\\cite{hcp1,hcp2,hcp3}, does not exist. The anisotropic expansion is achieved when the vector field has a VEV along one of the spatial directions. The anisotropic solution is an attractor, with anisotropy increasing towards the end of inflation, so these types of models are counter examples to the cosmic no hair conjecture~\\cite{Wald}. We have considered a generalization of the original model introduced in reference~\\cite{Watanabe:2009ct}, to include additional scalar fields, and specifically concentrated on the double scalar field case. This modification was done in order to achieve isotropization of the universe before the end of inflation. The original model with a single scalar field coupled to the vector field has an anisotropic inflationary background which persists until the end of inflation. In this type of background, perturbations at all scales are modified. In the two field modification, the extra field is not coupled to the vector field, and it is responsible for the overall isotropic expansion of the universe only. We have chosen the ratio of the masses of the scalar fields such that the inflationary expansion takes place in two steps, the first being anisotropic and the second is isotropic. Therefore, only the largest scale perturbations are modified and at small scales the standard spectrum is recovered. As we have discussed in the introduction, vector fields with a nonvanishing VEV have been introduced since isotropization takes place very quickly in a Bianchi-I background, (when anisotropy is only coming from initial conditions) and thus leads to a fine tuning. We have introduced a two phase inflation, with the second phase being isotropic. The transition to the second isotropic phase is indeed very quick, however the fine tuning has been relaxed compared to the previous case. The second isotropic phase should still be tuned to last around $60$ e-folds, but this tuning is not so strict, since it only affects the scale where the power spectrum becomes statistically isotropic. Previously, the tuning was more stringent, since inflation must last for $60$ e-folds in order that the largest scales are modified from the initial anisotropic conditions. A more important drawback of the inlfationary setup with only initial anisotropic conditions is that the initial singularity is very close to the time of isotropization, which is around one Hubble time. Thus, a calculation of the power spectrum shows that it becomes divergent at the largest scales~\\cite{gcp2}~\\footnote{This does not mean that the model is unstable, it reflects that nonlinear effects become important, and one has to go beyond linear perturbation theory.}. We have studied the perturbations around the background configuration in detail, taking into account all the degrees of freedom of the system. Following the formalism developed in references~\\cite{gcp,gcp2}, we have classified decoupled sets of perturbations around the anisotropic background and studied them separately. Furthermore, we have computed the quadratic action for perturbations and showed that the model is stable and consistent. We also found linear combinations of the perturbations which canonically normalize the action. This enabled us to quantize the system which in turn determined the initial conditions for the perturbations in the adiabatic vacuum. We then numerically integrated the equations of motion for the $2d$ vector perturbations for a range of comoving momenta, which resulted in the power spectrum of the gravitational wave polarization $h_{\\times}$. We have only computed the quadratic action and determined the canonical modes for the $2d$ scalar perturbations. We leave the study of the power spectra and the consequent phenomenology, which in turn is related to the curvature and $h_+$ gravitational wave spectrum, to a later publication. The computed spectrum for $h_{\\times}$ has certain interesting features, which might have some relevance to the observed CMB anomalies. First of all, the spectrum has angular dependence for $kk_{iso}$, so the higher multipoles are not affected from the vector field. We have shown that the power spectrum is formally of the form $P_{\\times}(\\overrightarrow{k}) = P_{\\times,\\, iso}(k) + \\delta P_{\\times}(k,\\xi)$. We expect that this will be true also for the spectrum of scalar modes. We have also provided a functional form for $\\delta P_{\\times}$ by fitting to the numerical spectrum in equation (\\ref{powerexp}). Phenomenologically acceptable modifications to the large scale spectrum should satisfy $\\vert \\delta P \\vert \\ll \\vert P_{iso} \\vert$, however, we have shown that for the model under discussion, this is not the case. Although our results are obtained for the gravitational wave mode $h_{\\times}$, we expect a similar behavior also for the curvature spectrum, which can be compared with observations. At larger scales, the power spectrum is greatly enhanced satisfying $\\vert \\delta P_{\\times} \\vert > \\vert P_{\\times, \\, iso} \\vert$. Although the anisotropy in the background is small (in the level of a percent, as can be seen from equation(\\ref{sigmadot})), the modification to the spectrum at large scales is not small. This is due to the fact that the background attractor solution is disconnected from the standard isotropic solution and therefore, the behavior of perturbations are dramatically different. There is no parameter in the attractor solution that connects it continuously to the isotropic FRW limit. The anisotropy is always small and proportional to $1/c^2$ and there is no finite limit in $c$ which can continuously deform the attractor solution to the isotropic solution. Therefore, the power spectrum cannot be approximated as to be the isotropic piece plus a small correction from anisotropy. This issue unfortunately limits the phenomenology of the model. Even though the difficulties involved, anisotropic backgrounds still have potentially interesting phenomenological consequences. Off-diagonal correlations of the $a_{lm}$ coefficients of CMB temperature fluctuations arise in such backgrounds and can have relevance to the alignment of lowest multipoles. Moreover, at the largest scales, non standard scalar to tensor ratio is obtained, which can be tested by upcoming CMB experiments. The gravitational wave modes $h_{+}$ and $h_{\\times}$ behave differently at large scales, therefore a future detection of gravitational waves might be tested against the consequences of an early stage of anisotropic inflation. In this paper, we have provided a solid example of anisotropic inflationary calculation of perturbations and computed the spectrum of $h_{\\times}$ polarization of gravitational waves. Although the obtained spectrum has phenomenologically limited implications, we believe that our results are still relevant, since it is the first complete study of power spectrum in an anisotropic inflationary model driven by a vector field to our knowledge. In a further publication, we will study the spectrum of the curvature ${\\cal R}$ perturbation, which could then be compared with observational data." }, "0910/0910.3690_arXiv.txt": { "abstract": "We describe the current status of solar modelling and focus on the problems originated with the introduction of solar abundance determinations with low CNO abundance values. We use models computed with solar abundance compilations obtained during the last decade, including the newest published abundances by Asplund and collaborators. Results presented here make focus both on helioseismic properties and the models as well as in the neutrino fluxes predictions. We also discuss changes in radiative opacities to restore agreement between helioseismology, solar models, and solar abundances and show the effect of such modifications on solar neutrino fluxes. ", "introduction": "} Solar models are a corner stone of stellar astrophysics. The determination of the solar interior structure through helioseismology and, only a few years later, the discovery that neutrinos change flavor, gave spectacular confirmations of our ability to model the Sun and, by extension, of other stars. However, a series of works starting with a redetermination of the photospheric oxygen solar abundance \\citep{allende} and finishing with a complete revision of solar abundances \\citep{ags05}, led to a strong reduction in the overall metallicity of the Sun driven by much lower CNO and Ne abundances than previously determined. Soon after, solar models that adopted the new composition were shown to have an interior structure at odds with helioseismology determinations. Since then, the so-called {\\em solar abundance problem} has been in the spotlight of solar (and stellar) astrophysics. As nicely put by \\citet{dp06}, it represents the incompatibility between the best solar atmosphere and interior models available. The effects of the low metallicity in the solar interior has been widely discussed in the literature. Among many other references, the reader can refer to \\citet{basu04,turck04,monta04,bs05,dp06} and \\citet{montecarlo}. Some attempts to constrain the solar metallicity independently of photospheric measurements can be found in \\citet{antia06,lin07} and \\citet{bisonii}. The connection between solar neutrinos and composition has also been discussed in different works, e.g. \\citet{turck04,bp04,bs05,montecarlo,bps08,wick}. Possible modifications in the physical inputs of solar models have also been discussed in connection to the {\\em solar abundance problem}. The reader can refer to \\citet{monta04,guzik05,dp06,castro07} just to mention some relevant works. In this article, we present a short review of the field and present new solar models that incorporate the most recent solar abundance determination by \\citet{agss09}. In \\S~\\ref{sec:models} we describe the main characteristics of the models used to obtain the core results presented here, including the different options for solar compositions we used. Results are presented in \\S~\\ref{sec:results} where helioseismology properties of the models and solar neutrino fluxes are discussed in the context of current observational and experimental data. In \\S~\\ref{sec:opac} we go to some length in discussing radiative opacities as a possible solution to the abundance problem, including the effects of opacities in solar neutrino fluxes. We summarize in \\S~\\ref{sec:conclu}. ", "conclusions": "} We have attempted, in this incomplete review, to describe the current status of the {\\em solar abundance problem} that originated with new determinations of solar photospheric abundances from Asplund and collaborators. Results discussed here are based on models computed with both the original \\citep{ags05} and the newest \\citep{agss09} solar compositions. Our reference for a {\\em good} solar model is based on the \\citet{gs98} composition. The most important result is that with the new \\citep{agss09} abundances, the qualitative picture that emerged a few years ago, i.e. that low-Z solar models are in gross disagreement with helioseismology, remains the same. Quantitatively, the disagreement is less severe because the new abundances have slightly higher CNO abundances and a somewhat larger Ne abundance. The changes, however, do not help much neither in restoring the agreement with helioseismology nor in facilitating the way for alternative solutions in the form of modified input physics for solar models. We have described with some detail the effect of the new composition in opacities and the required change to recover good helioseismic properties. Changes of order 15\\% are needed, which are still much higher than currently estimated uncertainties in radiative opacities for the solar interior. In addition to helioseismic properties of the models, we have discussed the effects of the composition on the predicted neutrino fluxes and compared, when possible, with results from solar neutrino experiments. Additionally, we have tried to encourage efforts to experimentally determine neutrino fluxes from CNO bicycle, since these are the most sensitive fluxes to changes in abundances of CNO elements, thus offering the best chances for neutrinos to put direct constraints on the solar core composition." }, "0910/0910.3373_arXiv.txt": { "abstract": "{ Multiple inflation is a model based on N=1 supergravity wherein there are sudden changes in the mass of the inflaton because it couples to `flat direction' scalar fields which undergo symmetry breaking phase transitions as the universe cools. The resulting brief violations of slow-roll evolution generate a non-gaussian signal which we find to be oscillatory and yielding $f_\\mathrm{NL} \\sim 5-20$. This is potentially detectable by e.g. Planck but would require new bispectrum estimators to do so. We also derive a model-independent result relating the period of oscillations of a phase transition during inflation to the period of oscillations in the primordial curvature perturbations generated by the inflaton. } ", "introduction": "A major goal in modern observational cosmology is to detect any non-gaussianity of the temperature anisotropies in the cosmic microwave background (CMB) \\cite{Komatsu:2001rj,Bartolo:2004if,Chen:2006nt}. All single field, slow roll, inflation models give rise to anisotropies that are gaussian \\cite{Maldacena:2002vr}. Therefore detection of a significant non-gaussian signal would falsify such models and provide new insights into the dynamics of inflation. Quantifying such a signal is however not a straightforward task. Whilst gaussianity is a well-defined property, non-gaussianity is not and the anisotropies can, in principle, deviate from gaussianity in many different ways \\cite{Chen:2006nt}. Also, the CMB temperature anisotropies measured by WMAP are \\emph{very} close to gaussian \\cite{Komatsu:2008hk}. Therefore any measure designed to detect non-gaussianity needs to be sensitive to a very small signal. One measure of non-gaussianity that is particularly useful is the three point correlation function, or `bispectrum' of the temperature anisotropies. Gaussian random variables (and functionals of gaussian random variables) have the property that all odd power correlation functions are zero. This makes the bispectrum the lowest order statistic for which \\emph{any} non-zero result would indicate a departure from gaussianity. The bispectrum contains much more information than the power spectrum as, in general, it depends on both scale and shape. Therefore, if a non-zero bispectrum is detected it will also be an extremely useful statistic for constraining models of the early universe. \\medskip However there are limitations to how much information can be inferred from the bispectrum: \\begin{itemize} \\item Modern CMB experiments possess very large numbers of pixels of data. e.g. $N \\simeq 10^6$ for WMAP and $N \\simeq 10^7$ for Planck \\cite{Ashdown:2006ey}. Considering that each higher order of correlation function will require an additional factor of $N$ calculations, exact determination of the bispectrum quickly becomes impossible. Therefore various estimators have been constructed \\cite{Fergusson:2008ra}, but each can look only for a specific type of bispectrum and so might miss a signal different from the type being searched for. \\item Higher order correlation functions suffer more from cosmic variance because more information is required at each scale to compute them. In this work we will deal with a \\emph{scale dependent} signal, therefore limitations due to cosmic variance are particularly relevant. \\end{itemize} These problems are not necessarily insurmountable. Computing power will only continue to get stronger, making bispectrum estimators increasingly powerful with time. For each potential signal, estimators can also be constructed for that specific signal, ensuring processing time is used as efficiently as possible. The estimator can be optimised for detecting primordial non-gaussianity while discriminating against secondary non-gaussianities arising from e.g. unsubtracted point sources or residuals from component separation \\cite{Munshi:2009ik}. Finally, in models that have non-trivial scale dependence, a correlation will likely exist between the power spectrum and bispectrum as we illustrate in this paper. One of the first methods for quantifying non-gaussianity involved rewriting the Newtonian potential $\\Phi(x)$ as \\cite{Komatsu:2001rj}, \\begin{equation} \\label{fNL local} \\Phi(x) = \\Phi_\\mathrm{L} (x) + f_\\mathrm{NL} (\\Phi_\\mathrm{L}^2 (x) - \\langle\\Phi_\\mathrm{L}^2 (x)\\rangle), \\end{equation} where $f_\\mathrm{NL}$ is a constant, and $\\Phi_\\mathrm{L} (x)$ is a gaussian variable. However, this parameterisation has a rather specific form and does not capture other possible deviations from gaussianity. A more general parameterisation can be obtained by allowing $f_\\mathrm{NL}$ to depend explicitly on the wavevector, ${\\bf k}$. In terms of the adiabatic curvature perturbation, $\\zeta$, the parameterisation becomes \\cite{Bartolo:2004if,Seery:2005wm}, \\begin{equation} \\label{fNL general} \\zeta(x) = \\zeta_\\mathrm{L} (x) - \\frac{3}{5} f_\\mathrm{NL}\\star(\\zeta_\\mathrm{L}^2 (x) - \\langle\\zeta_\\mathrm{L}^2 (x)\\rangle). \\end{equation} Here, the $\\star$ product is used because $f_\\mathrm{NL}$ has been allowed to depend on scale.\\footnote{See Eq.(229) in Ref.\\cite{Bartolo:2004if} for how to define this product.} The factor of -3/5 comes from the relationship between the adiabatic curvature and the Newtonian potential at matter domination. For $f_\\mathrm{NL}$ as defined in Eq.(\\ref{fNL local}), the WMAP 5-year constraint is $-9 < f_\\mathrm{NL} < 111$ \\cite{Komatsu:2008hk}. This reinforces the point made already: the primordial temperature anisotropies are \\emph{very} close to gaussian. Given that the amplitude of these anisotropies is $\\Delta T/T \\sim 10^{-5}$ the non-gaussian contribution to these anisotropies is at most $\\sim10^{-8}$ (or $\\sim 0.1\\%$ of the overall anisotropy). In general the expectation from inflation is that the anisotropies should be close to gaussian, therefore this result can be considered as a success of the inflationary paradigm. More precisely the expectation from inflation of non-gaussianity as parameterised by Eq.(\\ref{fNL general}) depends on the specific model. It is known that for single field, slow-roll inflation with a canonical kinetic term and a vacuum initial state, the result is $f_\\mathrm{NL} \\sim \\epsilon \\ll 1$ \\cite{Maldacena:2002vr} where $\\epsilon$ is the usual slow-roll parameter defined later in Eq.(\\ref{defeps}). The best sensitivity expected from the Planck satellite is to $f_\\mathrm{NL}$ of $\\mathcal{O}(5)$ while the secondary contribution from post-inflationary evolution is of $\\mathcal{O}(1)$ \\cite{Komatsu:2009kd}. Therefore the prediction of the simplest inflationary toy models~\\footnote{By this we mean a generic fine-tuned potential such as the frequently used $V (\\phi) = m^2 \\phi^2$ which has not yet been convincingly obtained from a physical theory, especially since inflation occurs in such models at $\\phi > M_\\mathrm{P}$.} is that there should \\emph{not} be a detection of primordial non-gaussianity in the near future. Conversely a deviation from the simplest toy models can produce a larger value for $f_\\mathrm{NL}$. The study of these effects serves two purposes. Firstly, from the cosmological perspective, a detection of $f_\\mathrm{NL}$ will immediately rule out single field slow-roll inflation and focus attention on determining how inflation actually occured. Secondly, from the perspective of inflationary model building based on fundamental physics, as the bounds on $f_\\mathrm{NL}$ grow tighter, some interesting models can be ruled out if there is no detection. Much work has been done on non-gaussianity generated due to multiple scalar fields (e.g. Ref.\\cite{Sasaki:2006kq}), non-canonical kinetic terms (e.g. Ref.\\cite{Langlois:2008qf}) or non-vacuum initial states (e.g. Ref.\\cite{Martin:1999fa}). However, surprisingly little attention has been paid to the possibility of non-gaussianity generated by a violation of slow-roll. Ref.\\cite{Byrnes:2009qy} did consider this in a context when multiple fields are present; however, the non-gaussianity itself is generated by the multiple fields, not the violation of slow-roll (see also Ref.\\cite{Cai:2009hw}). Refs.\\cite{Chen:2006xjb,Chen:2008wn} do consider the non-gaussianity generated by violation of slow-roll, but for a toy inflationary model. We follow their methods closely but consider a \\emph{physical} model of inflation. In both models the violation of slow-roll generates sharp features in the power spectrum and also generates a ringing in the bispectrum with a characteristic period which we calculate below. \\subsection{Features in the Power Spectrum of Primordial Fluctuations} It is an expectation in the simplest toy models of inflation that the power spectrum of temperature anisotropies in the CMB should be nearly scale-invariant \\cite{Lyth:1998xn}. However this need not be the case for physical models. If the observed spectrum was indeed perfectly scale-invariant, such models would be ruled out and na\\\"ively this might seem to be the case. Assuming the ``concordance'' $\\Lambda$CDM cosmology, the primordial spectrum is well parameterised as a power law with spectral index $n_\\mathrm{s} = 0.960 \\pm 0.014$ \\cite{Komatsu:2008hk}. This indicates that there cannot be any significant scale dependence that affects the spectrum over a large range of scales but it does \\emph{not} preclude the existence of e.g. localised oscillations in the spectrum, or other sharp features. Moreover since the observed anistrotopies arise from the convolution of the {\\em unknown} primordial spectrum with the transfer function of the {\\em assumed} cosmological model whose parameters are being determined, it is obvious that both unknowns cannot be determined simultaneously without further assumptions. In fact, there are indications that the primordial spectrum might not be a scale-free power law, even assuming the $\\Lambda$CDM cosmology \\cite{Martin:2003sg,Kogo:2004vt,Shafieloo:2003gf,Tocchini-Valentini:2004ht,Nicholson:2009pi,Ichiki:2009zz}. At large angular scales ($>50^0$) there is essentially no power and there are anomalous `glitches', especially in the range of multipoles $\\ell \\simeq 20-40$ \\cite{Spergel:2003cb,Hinshaw:2006ia}. The statistical significance of these anomalies is not sufficient to claim a definite detection, however it is strong enough to provide some tension with the fit to a scale-free power-law primordial spectrum which has only a 3\\% probability of being a good description of the WMAP 1-year data \\cite{Spergel:2003cb}, although this did improve to $\\sim7\\%$ with the WMAP 3-year data release \\cite{Hinshaw:2006ia}. Future measurements of the $EE$ and $TE$ mode polarisations by the Planck and (proposed) CMBPol satellites will throw light on whether these glitches are real or not \\cite{Mortonson:2009qv}. {\\it ``In the absence of an established theoretical framework in which to interpret these glitches (beyond the Gaussian, random phase paradigm), they will likely remain curiosities''} \\cite{Hinshaw:2006ia}. Indeed if there were a model that had predicted, without ambiguity, the position and amplitude of the glitches, this would be seen as very strong evidence for the model. Although no such model exists, the general possibility of generating glitches over a range of scales had in fact been proposed prior to the WMAP observations in the context of `multiple inflation' wherein the mass of the inflaton field undergoes sudden changes during inflation \\cite{Adams:1997de}. This generates characteristic localized oscillations in the spectrum, as was demonstrated numerically in a toy model of a inflationary potential with a `kink' parameterised as \\cite{Adams:2001vc}: \\begin{equation} V (\\phi) = \\frac{1}{2} m^2 \\phi^2 \\left[1 + c \\tanh \\left(\\frac{\\phi - \\phi_\\mathrm{s}}{d}\\right)\\right]. \\label{kinkpot} \\end{equation} By tuning the position, amplitude and gradient of the kink, the locations of the glitches can be varied to match the glitches seen in the power spectrum. One can then perform statistical likelihood tests to determine whether the fit is better with the glitches, but with the additional parameters, or with the simple scale-free spectrum. It was found by the WMAP team that the fit to the 1-year data improves significantly (by $\\Delta\\chi^2 = 10$) for the model parameters $\\phi_\\mathrm{s} = 15.5\\,M_\\mathrm{P}$, $c = 9.1\\times10^{-4}$ and $d = 1.4\\times10^{-2}\\,M_\\mathrm{P}$, where $M_\\mathrm{P}\\equiv(8\\pi\\,G_\\mathrm{N})^{-1/2} \\simeq 2.44\\times10^{18}$~GeV \\cite{Peiris:2003ff}. This analysis was repeated later using the WMAP 3-year data, with similar results \\cite{Covi:2006ci}. This seems encouraging, however $m$ in the toy model above is {\\em not} the mass of the inflaton --- in fact in all such monomial `chaotic' inflation models with $V \\propto \\phi^n$, inflation occurs at field values $\\phi_\\mathrm{infl} > M_\\mathrm{P}$, hence the leading term in a Taylor expansion of the potential around $\\phi_\\mathrm{infl}$ is always linear in $\\phi$ (since this is not a point of symmetry), rather than quadratic as for a mass term \\cite{German:1999gi}. The effect of a change in the inflaton mass can be sensibly modelled only in `new' inflation where inflation occurs at field values $\\phi_\\mathrm{infl} << M_{\\rm Pl}$ and an effective field theory description of the inflaton potential is possible. The `slow-roll' conditions are violated when the inflaton mass changes due to its (gravitational) coupling to `flat direction' fields which undergo thermal phase transitions as the universe cools during inflation \\cite{Adams:1997de}. The resulting effect on the spectrum of the curvature perturbation was found by analytic solution of the governing equations to correspond to a `step' followed by rapidly damped oscillations \\cite{Hunt:2004vt}.\\footnote{Although a similar phenomenon had been noted earlier for the case where the inflaton potential has a jump in its slope \\cite{Starobinsky:ts}, such a discontinuity has no physical interpretation.} The next step should be to predict other observables, having used the power spectrum to constrain all the parameters in the model. In this paper we calculate the bispectrum of the multiple inflation model \\cite{Adams:1997de}, using its predicted power spectrum \\cite{Hunt:2004vt} and the set of parameters which provide the best reduced $\\chi^2$ in the $\\Lambda$CDM cosmology \\cite{Hunt:2007dn}. We also examine the effect on the bispectrum of varying these parameters over their full natural range.\\footnote{We use the word `natural' in this context to mean ``stable towards radiative corrections''.} The non-gaussianity is found to be potentially detectable by the Planck satellite, or perhaps even a reanalysis of the WMAP data. ", "conclusions": "There is tentative evidence that the primordial power spectrum of scalar perturbations is imprinted with sharp features on large scales \\cite{Spergel:2003cb,Hinshaw:2006ia}. These could have resulted from one or more phase transitions early on in the inflationary epoch as happens in the multiple inflation model \\cite{Adams:1997de}. This model is well motivated by fundamental physics ($N=1$ supergravity) and such spectral features were predicted \\emph{before} WMAP provided observational indications for them. It was shown \\cite{Chen:2006xjb} that any such departure from slow-roll during inflation should also generate non-gaussianity. We have calculated the non-gaussianity in mutliple inflation for the parameter values that best fit the features in the power spectrum \\cite{Hunt:2007dn}. This is on the edge of being observable but a clear detection would require a new type of bispectrum estimator due to its scale dependence and non factorisability \\cite{Fergusson:2008ra}. Significantly larger non-gaussianities can be generated during multiple inflation if the parameters are allowed to range over values which are technically natural (i.e. stable towards radiative corrections). The form of this non-gaussianity is tightly correlated with the power spectrum --- the oscillations in the bispectrum should begin at the same multipole as the oscillations in the power spectrum and have two thirds of the period. There have been a number of attempts, to deconvolve the primordial power spectrum directly from the WMAP data \\cite{Kogo:2004vt,Shafieloo:2003gf,Tocchini-Valentini:2004ht} (see also Ref.\\cite{Bridges:2005br}). It is generally found that there is a supression of power on the scale of the present Hubble radius, followed by a `ringing' at medium scales. By measuring the period of the oscillations in the power spectrum one can predict the period (and phase) of oscillations in the bispectrum, in a \\emph{model independent} manner. Forthcoming measurements of CMB polarisation by Planck ought to shed light on whether these features in the power spectrum are systematic errors or genuine evidence of non-trivial dynamics during the inflationary era. The associated non-gaussian signal should also be detectable according to our model calculation and provide insight into the dynamics. \\subsection{Acknowledgements} SH acknowledges a Clarendon Fellowship and support from Balliol College, Oxford and the EU Marie Curie Network ``UniverseNet'' (HPRN-CT-2006-035863). We thank Paul Hunt, David Lyth, Graham Ross, Misao Sasaki and David Wands for helpful comments and discussions." }, "0910/0910.4765_arXiv.txt": { "abstract": "The next decade will bring massive new data sets from experiments of the direct detection of weakly interacting massive particle (WIMP) dark matter. Mapping the data sets to the particle-physics properties of dark matter is complicated not only by the considerable uncertainties in the dark-matter model, but by its poorly constrained local distribution function (the ``astrophysics'' of dark matter). I propose a shift in how to think about direct-detection data analysis. I show that by treating the astrophysical and particle-physics uncertainties of dark matter on equal footing, and by incorporating a combination of data sets into the analysis, one may recover both the particle physics and astrophysics of dark matter. Not only does such an approach yield more accurate estimates of dark-matter properties, but it may illuminate how dark matter coevolves with galaxies. ", "introduction": " ", "conclusions": "" }, "0910/0910.1971_arXiv.txt": { "abstract": "The \\amidas\\ website has been established as an online interactive tool for running simulations and analyzing data in direct Dark Matter detection experiments. At the first phase of the website building, only some commonly used WIMP velocity distribution functions and elastic nuclear form factors have been involved in the \\amidas\\ code. In order to let the options for velocity distribution as well as for nuclear form factors be more flexible, we have extended the \\amidas\\ code to be able to include {\\em user--uploaded} files with their own functions. In this article, I describe the preparation of files of user--defined functions onto the \\amidas\\ website. Some examples will also be given. ", "introduction": "In the last few years we developed new methods for analyzing data, i.e., measured recoil energies, from (future) direct Dark Matter detection experiments as model--independently as possible \\cite{DMDDf1v, DMDDmchi, DMDDfp2-IDM2008, DMDDidentification-DARK2009}. These methods will help us to understand the nature of WIMP (Weakly Interacting Massive Particle $\\chi$) Dark Matter, to identify them among new particles produced hopefully in the near future at colliders, as well as to reconstruct the (sub)structure of our Galactic halo. Following the development of these model--independent data analysis procedures, we combined the programs for simulations to a compact system: \\amidas\\ (A Model--Independent Data Analysis System). For users' convenience and under the collaboration with the ILIAS Project \\cite{ILIAS}, an online system has also been established at the same time \\cite{AMIDAS-web, AMIDAS-SUSY09}. For the first version of the \\amidas\\ code and website, the options for target nuclei, for the velocity distribution function of halo WIMPs, as well as for the elastic nuclear form factors for spin--independent (SI) and spin--dependent (SD) WIMP--nucleus interactions are fixed and only some commonly used forms have been involved in the \\amidas\\ code \\cite{AMIDAS-SUSY09}. Users can not choose different detector materials nor use different velocity distribution/form factors for their simulations and/or data analyses. In order to let the options for the velocity distribution as well as for the nuclear form factors be more flexible, we have extended the \\amidas\\ code to be able to include {\\em user--uploaded} files with their own functions. Note that, since the \\amidas\\ code has been written in the C programming language, all user--defined functions for uploading must also be given using the syntax of C. The remainder of this article is organized as follows. In Sec.~2 I will talk about setting users' own target nuclei. In Secs.~3 and 4 the preparation of files defining the velocity distribution function and the nuclear form factors will be described, respectively. I will conclude in Sec.~5. Some intrinsically defined constants and functions in the \\amidas\\ code will be given in an appendix. ", "conclusions": "In this article, I described the preparation of files giving user--defined WIMP velocity distribution function and/or elastic nuclear form factor(s) which can be uploaded onto the \\amidas\\ website for more flexible simulations or data analyses. This improvement allows theorists to simulate with their own/favorite models and compare them with (future) experimental results, as well as gives experimentalists flexible choices for more suitable form factor(s) for their own detector materials. In summary, up to now all basic functions of the \\amidas\\ code and website have been well established. Hopefully this new tool can help our colleagues to detect/discover WIMP Dark Matter, to understand the nature of Dark Matter particles and the (sub)structure of the Galactic halo in the future. \\subsubsection*" }, "0910/0910.4881_arXiv.txt": { "abstract": "We report on observations of TeV-selected AGN made during the first 5.5 months of observations with the Large Area Telescope (LAT) on-board the \\textit{Fermi Gamma-ray Space Telescope} (\\Fermic). In total, \\NObjTotal\\ AGN were selected for study, each being either (i) a source detected at TeV energies (\\NObjTeVSrc\\ sources) or (ii) an object that has been studied with TeV instruments and for which an upper-limit has been reported (\\NObjTeVLim\\ objects). The \\Fermi observations show clear detections of \\NDetTotal\\ of these TeV-selected objects, of which \\NDetTeVSrc\\ are joint GeV--TeV sources and \\NDetTotalNoEGRET\\ were not in the third EGRET catalog. For each of the \\NDetTotal\\ \\Fermi-detected sources, spectra and light curves are presented. Most can be described with a power law of spectral index harder than 2.0, with a spectral break generally required to accommodate the TeV measurements. Based on an extrapolation of the \\Fermi spectrum, we identify sources, not previously detected at TeV energies, which are promising targets for TeV instruments. Evidence for systematic evolution of the \\gray spectrum with redshift is presented and discussed in the context of interaction with the EBL. ", "introduction": " ", "conclusions": "" }, "0910/0910.2375_arXiv.txt": { "abstract": "Computer representations of real numbers are necessarily discrete, with some finite resolution, discreteness, quantization, or minimum representable difference. We perform astrometric and photometric measurements on stars and co-add multiple observations of faint sources to demonstrate that essentially all of the scientific information in an optical astronomical image can be preserved or transmitted when the minimum representable difference is a factor of two finer than the root-variance of the per-pixel noise. Adopting a representation this coarse reduces bandwidth for data acquisition, transmission, or storage, or permits better use of the system dynamic range, without sacrificing any information for down-stream data analysis, including information on sources fainter than the minimum representable difference itself. ", "introduction": "Computers operate on bits and collections of bits; the numbers stored by a computer are necessarily discrete; finite in both range and resolution. Computer-mediated measurements or quantitative observations of the world are therefore only approximately real-valued. This means that choices must be made, in the design of a computer instrument or a computational representation of data, about the range and resolution of represented numbers. In astronomy this limitation is keenly felt at the present day in optical imaging systems, where the analog-to-digital conversion of CCD or equivalent detector read-out happens in real time and is severely limited in bandwidth; often there are only eight bits per readout pixel. This is even more constrained in space missions, where it is not just the bandwidth of real-time electronics but the bandwidth of telemetry of data from space to ground that is limited. If the ``gain'' of the system is set too far in one direction, too much of the dynamic range is spent on noise, and bright sources saturate the representation too frequently. If the gain is set too far in the other direction, information is lost about faint sources. Fortunately, the information content of any astronomical image is limited \\emph{naturally} by the fact that the image contains \\emph{noise}. That is, tiny differences between pixel values---differences much smaller than the amplitude of any additive noise---do not carry very much astronomical information. For this reason, the discreteness of computer representations of pixel values do not have to limit the scientific information content in a computer-recorded image. All that is required is that the noise in the image be \\emph{resolved} by the representation. What this means, quantitatively, for the design of imaging systems is the subject of this \\documentname; we are asking this question: ``What bandwidth is required to deliver the scientific information content of a computer-recorded image?'' This question has been asked before, using information theory, in the context of telemetry \\citep{Gaztanaga} or image compression \\citep{Watson}, treating the pixels (or linear combinations of them) as independent. Here we ask this question, in some sense, \\emph{experimentally}, and for the properties of imaging on which optical astronomy depends, where groups of contiguous pixels are used in concert to detect and centroid faint sources. We perform experiments with artificial data, varying the bandwidth of the representation---the size of the smallest representable difference $\\Delta$ in pixel values---and measuring properties of scientific interest in the image. We go beyond previous experiments of this kind (\\citealt{WhiteGreenfield}, and \\citealt{PenceInPress}) by measuring the centroids and brightnesses of compact sources, and sources fainter than the detection limit. The higher the bandwidth, the better these measurements become, in precision and in accuracy. We find, in agreement with previous experiments and information-theory-based results, that the smallest representable difference $\\Delta$ should be on the order of the root-variance $\\sigma$ of the noise in the image. More specifically, we find that \\emph{the minimum representable difference should be about half the per-pixel noise sigma} or that about two bits should span the FWHM of the noise distribution if the computer representation is to deliver the information content of the image. Of course, tiny mean differences in pixel values, even differences much smaller than the noise amplitude, \\emph{do} contain \\emph{extremely valuable} information, as is clear when many short exposures (for example) of one patch of the sky are co-added or analyzed simultaneously. ``Blank'' or noise-dominated parts of the individual images become signal-dominated in the co-added image. In what follows, we explicitly include this ``below-the-noise'' information as part of the information content of the image. Perhaps surprisingly, \\emph{all} of the information can be preserved, even about sources fainter than the discreteness of the computer representation, provided that the discreteness is finer than the amplitude of the noise. This result has important implications for image compression, but our main interest here is in the design and configuration of systems that efficiently take or store raw data, using as much of the necessarily limited dynamic range on signal as possible. Our results have some relationship to the study of \\emph{stochastic resonance}, where it has been shown that signals of low dynamic range can be better detected in the presence of noise than in the absence of noise (see \\citealt{stochres} for a review). These studies show that if a signal is below the minimum representable difference $\\Delta$, it is visible in the data only when the digitization of the signal is noisy. A crude summary of this literature is that the optimal noise amplitude is comparable to the minimum representable difference $\\Delta$. We turn the stochastic resonance problem on its head: The counterintuitive result (in the stochastic resonance context) that weak signals become detectable only when the digitization is noisy becomes the relatively obvious result (in our context) that so long as the minimum representable difference $\\Delta$ is comparable to or smaller than the noise, signals are transmitted at the maximum fidelity possible in the data set. What we call here the ``minimum representable difference'' has also been called by other authors the ``discretization'' \\citep{Gaztanaga} or ``quantization'' \\citep{Watson, WhiteGreenfield, Pence}. ", "conclusions": "Because of finite noise, the information content in astronomical images is finite, and can be captured by a finite numerical resolution. In the above, we scaled and snapped-to-integer real-valued images by a SNIP procedure such that in the SNIPped image, the minimum representable difference $\\Delta$ between pixel values was set to a definite fraction of the Gaussian noise root-variance (sigma) $\\sigma$. We found with direct numerical experiments that the SNIP procedure introduces essentially no significant error in estimating the variance of the image, or in centroiding or photometering stars in the image, when the minimum representable difference is set to any value $\\Delta \\leq 0.5\\,\\sigma$. In addition, we showed that all the information about sources fainter than the per-pixel noise level is preserved by the quantization (SNIP) procedure, again provided that $\\Delta \\leq 0.5\\,\\sigma$. This is somewhat remarkable because at $\\Delta = 0.5\\,\\sigma$ the faintest sources in our experiments were fainter than the minimum representable difference. Although it is somewhat counterintuitive that integer quantization of the data does not remove information about sources fainter than the quantization level, it is perhaps even more counterintuitive how well photometric measurements perform in our coadd tests. For example, in \\figurename s~\\ref{fig:bitsoffsetcoadd1} and \\ref{fig:bitsoffsetcoadd2}, the photometric measurements are relatively accurate even when the data are quantized at minimum representable difference $\\Delta = 16\\,\\sigma$! The quality of the measurements can be understood in part by noting that the coadded images have a per-pixel noise $\\sigma = \\sqrt{1024}\\,\\sigma = 32\\,\\sigma$, which is once again larger than the minimum representable difference, and in part by noting that each image has a different sky level, so each individual image is differently ``wrong'' in its photometry; many of these differences average out in the coadd. When the sky level is held fixed across coadded images, photometric measurements become inaccurate again---as seen in \\figurename~\\ref{fig:bitsoffsetcoadd1_samesky}---because individual-image biases caused by the coarse quantization no longer ``average out''. Our fundamental conclusion is that all of the scientifically relevant information in an astronomical image is preserved as long as the minimum representable difference $\\Delta$ in pixel values is smaller than or equal to half the per-pixel root-variance (sigma) $\\sigma$ in the image noise. This confirms previous results based on information-theory arguments (for example, \\citealt{Gaztanaga}), and extends previous experiments on bright-source photometry \\citep{WhiteGreenfield, PenceInPress} to astrometry and to sources fainter than the noise. Our experiments were performed on images with pure Gaussian noise; of course many images contain significant non-Gaussianity in their per-pixel noise so the empirical variance will depart significantly from the true noise variance \\citep{WhiteGreenfield}. The conservative approach for such images is to take not the true variance for $\\sigma^2$ but rather use for $\\sigma^2$ something like the minimum of the straightforwardly measured variance and a central variance estimate, such as a sigma-clipped variance estimate, an estimate based on the curvature of the central part of the noise value frequency distribution function, or the median absolute difference of nearby pixels \\citep{Pence}. With this re-definition of the root variance $\\sigma$, the condition $\\Delta \\leq 0.5\\,\\sigma$ represents a conservative setting of the minimum representable difference. The fact that a $\\Delta = 0.5\\,\\sigma$ representation preserves information on the faint sources---even those fainter than $\\Delta$ itself---has implications for the design of data-taking systems, which are necessarily limited in bandwidth. If the system is set with $\\Delta$ substantially smaller than $0.5\\,\\sigma$, then bright sources will saturate the representation more frequently than necessary, while no additional information is being carried about the faintest sources. Any increase in $\\Delta$ pays off directly in putting more of the necessarily limited system dynamic range onto bright sources, so it behooves system designers to push as close to the $\\Delta = 0.5\\,\\sigma$ limit as possible. To put this in the context of a real data system, we looked at a ``DARK'' calibration image from the Hubble Space Telescope Advanced Camera for Surveys (ACS). The dark image should have the lowest per-pixel noise of any ACS image, because it has only dark and read noise. We chose image set \\texttt{jbanbea2q}, and measured the median noise level in the raw DARK image with the median absolute difference between values of nearby pixels (for robustness). The ACS data system is operating with a minimum representable difference $0.25\\,\\sigma < \\Delta < 0.33\\,\\sigma$, comfortably within the information-preserving range and close to the minimum-bandwidth limit of $\\Delta = 0.5\\,\\sigma$. Of course this is for a dark frame; sky exposures (especially long ones) could have been profitably taken with a larger $\\Delta$ (because $\\sigma$ will be greater); this would have preserved more of the system dynamic range for bright sources. If the ACS took almost exclusively long exposures, the output would contain more scientific information with a larger setting of the minimum representable difference. In some sense, the results of this paper recommend a ``lossy'' image compression technique, in which data are scaled by a factor and snapped to integer values such that the minimum representable difference $\\Delta$ is made equal to or smaller than $0.5\\,\\sigma$. Indeed, when typical real-valued astronomical images are converted to integers at this resolution, the integer versions compress far better with subsequent standard file compression techniques (such as gzip) than do the floating-point originals \\citep{Gaztanaga, Watson, WhiteGreenfield, Pence, bernstein}. In the $\\Delta = 0.5\\,\\sigma$ representation, after lossless compression, storage and transmission of the image ``costs'' only a few bits per noise-dominated pixel. Because the snap-to-integer step changes the data, this overall procedure is technically lossy, but we have shown here that none of the \\emph{scientific} information in the image has been lost." }, "0910/0910.5280_arXiv.txt": { "abstract": "This is my contribution to Proceedings of the International Workshop on Cosmic Structure and Evolution, September 23-25, 2009, Bielefeld , Germany. In my talk I presented some non-Gaussian features of the foreground reduced WMAP five year full sky temperature maps, which were recently reported in the Ref. \\cite{Vitaly}. And in these notes I first discuss the statistics behind this analysis in some detail. Then I describe invaluable insights which I got from discussions after my talk on the Workshop. And finally I explain why, in my current opinion, the signal detected in the Ref. \\cite{Vitaly} can hardly have something to do with cosmological perturbations, but rather it presents a fancy measurement of the Milky Way angular width in the microwave frequency range. ", "introduction": "Nowadays we witness a great progress in both theoretical and observational cosmology which makes our demands and expectations ever higher and turns us to discussing more and more subtle properties of the available data. One of such popular topics is the quest for primordial non-Gaussianities in the spectrum of the CMB radiation. Not really expected to be detectable for the simplest models of inflation, these small departures from the purely Gaussian signal would help to distinguish between more elaborate inflationary scenarios and would probably provide us with some new insights into the wonderful realm of the very early Universe. The approach I discuss is based on a very simple idea. Assume that the Universe is statistically isotropic, and all the temperature fluctuations in the CMB radiation are of statistical nature. Then we decompose the fluctuations, as usual, into the spherical harmonics denoting the coefficients by $a_{l,m}$ and get $$\\langle a_{l,m}a_{l^{\\prime},m^{\\prime}}^{*}\\rangle=\\langle a_{l,m}a_{l^{\\prime},-m^{\\prime}}\\rangle =C_l\\delta_{l,l^{\\prime}}\\delta_{m,m^{\\prime}}.$$ Moreover, $a_{l,m}$'s with a fixed value of $l$ but different values of $m$ can be thought of as different realizations of one and the same random variable. So that one can naturally ask a question about the shape of the distribution. It can be answered by many methods, from plotting a histogram of the sample to estimating the higher moments of the distribution. A Gaussian distribution is completely determined by two parameters, its mean and its variance. For fluctuations the mean is taken to be $0$, and the only parameter left is the variance, $C_l$. (Of course, it implies rescaled $\\chi^2$-distributions for quadratic in $a$ quantities.) If the random variable is known (or assumed) to be Gaussian, this single parameter can, in principle, be extracted from any part of the probability distribution. The idea of the Ref. \\cite{Vitaly} is to take only the tails, i.e. to deduce the variance $C_l$ once more from only the distribution of large coefficients, $|a_{l,m}|^2>C_l$, and then to compare this result with the original one. Up to statistical variations in the number of data points in the tails, this corresponds to using the order statistics with somewhat more points than in top and bottom sextiles but with less points than in two marginal quintiles. Applied to the full sky foreground reduced maps, this method gives a well-pronounced peak in the difference between the two estimates for the variance. The peak is located in the range of $l\\approx 45 \\pm 15$. The fluctuations outside of the peak are also much larger than would be expected statistically. This is, of course, due to remaining foregrounds contamination, and the only reason to take the peak seriously was that it is a few more times larger than the other fluctuations \\cite{Vitaly} and it looks more or less the same in different frequency bands (up to the different overall level of noise). And this was also my conclusion that it should have something to do with cosmology. I explain the relevant statistics in Sections 2 and 3. And for the graphical presentation of results, I refer the reader to \\cite{Vitaly}. However, at the Workshop I have learned from Pavel Naselsky that the multipole coefficients with even values of $l+m$ have the worst contamination from Galaxy, see below. In Section 4 I discuss the data analysis with separation of $(l+m)$-even and $(l+m)$-odd harmonics, and show that the initial assumption of having different realizations of one random variable is heavily disproved due to Galactic signal which invalidates the claim for cosmological non-Gaussianities. My current conclusion presented in the Section 5 is that the effects observed in \\cite{Vitaly} refer to the structure of Galactic noise, and not to properties of primordial fluctuations. Due to this reason I cancelled my authorship for the second version of that article. (I was a co-author for the first one). And I would like to note here that all the computer work for the article \\cite{Vitaly} was done by my former co-author Vitaly Vanchurin and, needless to say, if there appears to be something primordial about this peak then the whole success should be attributed solely to him and his enthusiasm. An interested reader may also want to consult with the original reference \\cite{Vitaly} for the opinion opposite to mine. ", "conclusions": "In these notes I discussed the statistics behind the non-Gaussian anomalies reported in Ref. \\cite{Vitaly}. After that I have shown that the most probable explanation of the signal refers to geometric properties of the Galactic noise, and not to cosmology. Admittedly, I do not have a good understanding of the structure of Galactic noise. But at the very least, the claim for cosmological non-Gaussianities is pretty much premature. (I refer an interested reader to the work \\cite{Vitaly} for a different opinion.) On the other hand, one could probably use this kind of analysis to extract some information about the structure of the noise. {\\bf Acknowledgments.} I am very grateful to the organizers of the Workshop for the opportunity to participate in this wonderful event, and especially I wish to thank Dominik Schwarz for invitation and for his encouragement to write this contribution. I am also very grateful to Pavel Naselsky and to other participants for very useful discussions. This work was supported in part by the Cluster of Excellence EXC 153 {}``Origin and Structure of the Universe''." }, "0910/0910.0831_arXiv.txt": { "abstract": "We report the first interferomteric detection of 183 GHz water emission in the low-mass protostar Serpens SMM1 using the Submillimeter Array with a resolution of 3$''$ and rms of $\\sim$7 Jy in a 3 km s$^{-1}$ bin. Due to the small size and high brightnessof more than 240 Jy/beam, it appears to be maser emission. In total three maser spots were detected out to $\\sim$ 700 AU from the central protostar, lying along the red-shifted outflow axis, outside the circumstellar disk but within the envelope region as evidenced by the continuum measurements. Two of the maser spots appear to be blue-shifted by about 1 to 2 km s$^{-1}$. No extended or compact thermal emission from a passively heated protostellar envelope was detected with a limit of 7 Jy (16 K), in agreement with recent modelling efforts. We propose that the maser spots originate within the cavity walls due to the interaction of the outflow jet with the surrounding protostellar envelope. Hydrodynamical models predict that such regions can be dense and warm enough to invert the 183 GHz water transition. ", "introduction": "Water is one of the most important molecules in interstellar clouds in general and in star-forming regions in particular. In warm regions ($T>$100 K) close to the protostar, water is prominent with gas abundances up to 3$\\times10^{-4}$ w.r.t. \\HH, even higher than CO \\citep{Cernicharo90,vanDishoeck96,Harwit98}. Besides the inner regions of protostellar envelopes these high abundances are also found where powerful jets from the protostar interact with the surroundings \\citep{Nisini99}. In contrast, water abundances are as low as $10^{-8}$-$10^{-9}$ in cold ($T<$100 K) envelope regions as evidenced by Infrared Space Observatory (ISO) and Submillimeter Wave Astronomy Satellite (SWAS) observations \\citep[e.g][]{Boonman03}. Emission from water molecules is very difficult to observe, due to the limits imposed by the Earth's atmosphere. Isotopologues such as deuterated water can be observed from the ground \\citep[e.g.][]{Schulz91,Parise05} The only transitions of the main water isotope that can be observed regularly are low-frequency maser transitions. Most famous is the water maser at 22.2 GHz, the 6$_{16}$-5$_{23}$ transition, regularly observed using radio telescopes and interferometers, such as the Very Large Array (VLA). In star forming regions with embedded sources of high luminosity ($>$100 L$_{\\odot}$) 22.2 GHz water maser emission is commonly detected and used to probe gas kinematics \\citep[e.g.][]{Moscadelli06,Goddi06}. The emission is found to be variable on timescales of a day to a month, probably related to variations in the accretion disk or the outflowing material \\citep[e.g.][]{Pashchenko05}. In regions with embedded sources of lower luminosity, 22.2 GHz water maser emission is detected less frequently than in high-mass sources, but again with considerable variability \\citep[e.g.][]{Wilking94,Claussen96,Claussen98,Furuya03}. Imaging with the VLA detected the maser emission within several hundred AU of low-mass protostars \\citep[e.g][]{Furuya99,Furuya01}, while Very Long Baseline Array observations indicate that the location may even be closer to the protostar \\citep[e.g][]{Moscadelli06}. Both water and methanol masers have been observed to be related to disks \\citep{Torrelles96,Moscadelli06}, while other observations of masers have been associated with outflows \\citep[e.g.][]{Claussen96,Furuya03}. However, the excitation conditions for the 22 GHz water maser line are very high density ($>$10$^8$ cm$^{-3}$) and temperature (2000 K$>$ T $>$200 K)\\citep{Yates97}, a combination of conditions that is rare in low-mass protostars. Another water maser transition is found at 183.3 GHz. This 3$_{13}$-2$_{20}$ transition has been used in extra-galactic and galactic studies, primarily with the 30m telescope at Pico Veleta, Spain \\citep[e.g.][]{Cernicharo94,Cernicharo96}. Statistical equilibrium calculations combined with the observation of extended emission conclude that relatively low temperatures (T$\\sim$ 150 K) and densities (10$^5$-10$^6$ cm$^{-3}$) can already invert the populations of the 183.3 GHz transition \\citep{Cernicharo94}. In star forming regions it has been detected in several low- and high-mass protostellar sources such as Orion, W49N, H7-11 and L1448-mm. Most lines consist of a broad component, superposed with a strong narrow line, presumably a maser line. The broad component was found to be spatially extended thermal emission after having been mapped in some high-mass star forming regions such as Orion and W49N \\citep{Cernicharo94,GonzalesAlfonso95}. Extended thermal emission was also found in low-mass sources such as HH7-11 \\citep{Cernicharo96}. Serpens SMM1 (referred to as SMM1) is a low-mass Class 0 source in the Serpens cluster ($D$= 250 pc) and has been studied at (sub)millimeter wavelengths by \\citet{Hogerheijde99}. This source was selected as part of a pilot study to observe the 183 GHz line due to its inclusion in the Water in Star-forming regions with Herschel (WISH) program\\footnote{See http://www.strw.leidenuniv.nl/WISH} as a source where many water lines will be targetted with spatial resolutions of 9-40$''$. It is relatively luminous ($L$=20.7 $L_\\odot$) and thus an ideal target for water observations. It was observed with ISO-LWS \\citep{Larsson02}, where numerous thermal water lines were found in the large 120$''$ beam. It drives a powerful highly collimated radio jet \\citep{Curiel93}, a large molecular outflow \\citep{White95} and was included in the 22 GHz water maser survey by \\citet{Furuya03}, where several maser spots were found. These masers were in turn studied by \\citet{Moscadelli06} who found the masers originating within a circumstellar disk based on their position and proper motions. Observations with SWAS (beam = 3.3$'$ by 4.5$'$) of the 1$_{10}$-1$_{01}$ water line found water associated with the outflow with an abundance of o-H$_2$O of a few times 10$^{-7}$ \\citep{Franklin08} in both red and blue outflow lobes. In this letter, we present first interferometric observations using the SMA\\footnote{The Submillimeter Array is a joint project between the Smithsonian Astrophysical Observatory and the Academia Sinica Institute of Astronomy and Astrophysics and is funded by the Smithsonian Institution and the Academia Sinica.} of the 183 GHz maser emission from a region of low mass star formation, associated with the protostar SMM1. % ", "conclusions": "The conclusions of this letter are as follow \\begin{itemize} \\item The 183 GHz emission of Serpens SMM1 is masing from three distinct maser spots at 500-1200 AU, aligned in the direction of the red-shifted outflow, two of which are slightly blue-shifted. \\item The extended dust continuum emission detected with the SMA in Serpens SMM1 comes from a protostellar envelope with an estimated mass of 2.7 M$_\\odot$. \\item A resolved component detected on baselines up to $\\sim$400 k$\\lambda$ indicates the presence of disk with an estimated mass of 0.1 M$_\\odot$. \\item It is theorized that the maser spots originate in Kelvin-Helmholtz instabilities in the red-shifted outflow, created by the jet interactions with the envelope. Such interactions would temporarily create small regions of high density and temperature, providing the conditions for the 183 GHz transition to mase in low-mass protostellar environment. \\item No compact or extended thermal water emission is detected, in agreement with a model of a passively heated envelope. \\end{itemize} Research into the 183 GHz maser is promising as an additional constraint to jet mechanics and interaction of the jet with its surrounding material, both the envelope and the outflow. In the next few years, the receivers of the SMA provide a unique opportunity to study the water maser emission at this wavelength in many low-mass YSOs. In the long run, ALMA will be able to observe these lines at higher sensitivity and resolution, possibly probing the scales at which these masers emit. Combined with high-resolution VLA and VLBI observations of the 22 GHz maser, such observations can truly probe the physical structure of maser-emitting regions. \\ack TvK is supported as an SMA postdoctoral fellow. Steve Longmore and Jes J{\\o}rgensen are thanked for discussion on fitting envelope models on continuum, and Elizabeth Humphreys for extensive discussion on maser excitation. TvK is also grateful to Lars Kristensen for providing the DUSTY model. The extensive ongoing discussions with Ewine van Dishoeck on many aspects of interstellar water are much appreciated. \\begin{figure}[!ht] \\begin{center} \\includegraphics[width=400pt]{./figure1.ps} \\end{center} \\caption{Moment map of the three maser spots in SMM1. Maser spots are labeled with their respective number. An X at the center 0,0 position marks the maximum intensity of the continuum (see Fig. \\ref{1:cont}), which coincides with the location of the protostar. The spectra at the three maser spots are shown . Due to the proximity of the spot 2 to both spot 1 and 3 ($\\sim$ 1 beamsize), the spectrum there is a blend of all three maser spots. The beamsize is to 2.7$''$ by 4.0$''$ with a P.A. of 49$^\\circ$.} \\label{1:moment} \\end{figure} \\begin{figure}[!ht] \\begin{center} \\includegraphics[width=220pt]{./figure2a.ps} \\includegraphics[width=220pt]{./figure2b.ps} \\includegraphics[width=230pt]{./figure2c.ps} \\includegraphics[width=230pt]{./figure2d.ps} \\end{center} \\caption{{\\it TOP:} The continuum map of SMM1 at 187 GHz. Levels are in 2$\\sigma$, 4$\\sigma$, 6$\\sigma$,... with $\\sigma$=15 mJy/beam. {\\it BOTTOM:} The UV distance versus amplitude plot, with the mean visibility amplitude in that annulus on the vertical axis and baseline length on the horizontal. The triangles show the observed visibilities of SMM1 at 187 GHz. Errors shown are the standard deviation in the mean, while the solid line shows the expectation value assuming no signal. The dashed line shows the expected visibilities for an envelope model at 194 GHz (Kristensen et al. in prep). The crosses represent the residual visibilities if the envelope model is subtracted from the observed visibilities.} \\label{1:cont} \\end{figure} \\begin{table}[!ht] \\caption{Properties of the maser spots} \\begin{center} \\begin{tabular}{l l l l l l} \\hline \\hline Maser spot & FWHM & $I_{\\rm{peak}}$ & $T_{\\rm{B}}$$^a$ &$V_{\\rm{LSR}}$ & Spat. Offset \\\\ & km s$^{-1}$ & Jy/beam & K &km s$^{-1}$ & RA,Dec ('','') \\\\ \\hline 1 & 0.7 & 463& 1871 & 9.2 & -3.0,0.2 \\\\ 2 & 1.5 & 243 & 982 &7.4 & -4.5,2.2\\\\ 3 & 1.3 & 328 & 1325 & 6.3 & -6.7,3.8\\\\ \\hline \\end{tabular} \\end{center} $^a$ assuming the emission fills the beam \\label{1:tab} \\end{table}" }, "0910/0910.5625_arXiv.txt": { "abstract": "The decay of dark matter is predicted by many theoretical models and can produce observable contributions to the cosmic-ray fluxes. I shortly discuss the interpretation of the positron and electron excess as observed by PAMELA and Fermi LAT in terms of decaying dark matter, and I point out the implications for the Fermi LAT observations of the $\\gamma$-ray flux with emphasis on its dipole-like anisotropy. ", "introduction": "The most popular type of dark matter (DM) candidate, the weakly interacting massive particle (WIMP), can naturally reproduce the observed DM abundance due to effective self-annihilation in the early Universe, and today this same annihilation process could produce an observable contribution to the measured cosmic-ray fluxes on Earth. Such an indirect detection of DM is also possible if DM \\textit{decays} with a sufficiently large rate. There exist a number of interesting DM models (see \\textit{e.g.}~\\cite{models} and references therein) that predict the decay of DM on cosmological time scales, namely with lifetimes around and above $\\tau_\\text{DM}\\simeq \\mathcal{O}(10^{26}\\,\\text{s})$, which are typically required to be not in conflict with current observational limits. Among these models is the gravitino with a small violation of $R$-parity, motivated by requiring a consistent thermal history of the Universe, and the sterile neutrinos, whose long lifetime is due to tiny Yukawa couplings. The typical masses for these DM candidates lie in the $100\\,\\text{GeV}$ and the $10\\,\\text{keV}$ regime, respectively. Another interesting model with kinetically mixed hidden gauginos was also recently studied~\\cite{hg}. Even in models where DM is stable in the first place, the consideration of higher-dimensional operators often renders the DM particle unstable with cosmological lifetimes. Since the indirect detection signals from decay differ in general from the ones of annihilation, a dedicated study of decaying DM signals is mandatory. Below I will shortly review the $\\gamma$-ray and $e^\\pm$-signals that can come from DM decay, and I will discuss them in light of recent observations. ", "conclusions": "Many theoretical models predict the decay of DM on cosmological timescales, giving rise to an anomalous contribution to the observed cosmic-ray fluxes. The corresponding $\\gamma$-ray signals could show up as broad features over large angular distance in the $\\gamma$-ray sky. If decaying DM is the right explanation of the positron and electron excess observed by PAMELA and Fermi LAT, a corresponding $\\gamma$-ray signal with a large dipole-like anisotropy should be observed in the very near future with Fermi LAT. This anisotropy would be due to prompt radiation at high latitudes, and due to ICS radiation at lower latitudes, most prominent in a region of a few kpc around the galactic center. It is tempting to speculate that such an ICS signal already showed up in the Fermi LAT data, see Ref.~\\cite{Dobler:2009xz}." }, "0910/0910.2233_arXiv.txt": { "abstract": "We have built a reliable and robust system that takes as input an astronomical image, and returns as output the pointing, scale, and orientation of that image (the astrometric calibration or WCS information). The system requires no first guess, and works with the information in the image pixels alone; that is, the problem is a generalization of the ``lost in space'' problem in which nothing---not even the image scale---is known. After robust source detection is performed in the input image, asterisms (sets of four or five stars) are geometrically hashed and compared to pre-indexed hashes to generate hypotheses about the astrometric calibration. A hypothesis is only accepted as true if it passes a Bayesian decision theory test against a null hypothesis. With indices built from the USNO-B Catalog and designed for uniformity of coverage and redundancy, the success rate is $>99.9~\\percent$ for contemporary near-ultraviolet and visual imaging survey data, with no false positives. The failure rate is consistent with the incompleteness of the USNO-B Catalog; augmentation with indices built from the 2MASS Catalog brings the completeness to $100~\\percent$ with no false positives. We are using this system to generate consistent and standards-compliant meta-data for digital and digitized imaging from plate repositories, automated observatories, individual scientific investigators, and hobbyists. This is the first step in a program of making it possible to trust calibration meta-data for astronomical data of arbitrary provenance. ", "introduction": " ", "conclusions": "" }, "0910/0910.2227_arXiv.txt": { "abstract": "We select 25,000 galaxies from the NEWFIRM Medium Band Survey (NMBS) to study the rest-frame $U~-~V$ color distribution of galaxies at $01$. There is some evidence that the color bimodality persists up to $z=2$, though the contrast separating the red and blue sequences is low. This is either because spectroscopic samples that span a broad range of colors are small \\markcite{giallongo:05,franzetti:07, cassata:08, kriek:08}({Giallongo} {et~al.} 2005; {Franzetti} {et~al.} 2007; {Cassata} {et~al.} 2008; {Kriek} {et~al.} 2008) or because larger photometric samples suffer from uncertain photometric redshifts at $z>1.5$, which leads to uncertain rest-frame colors \\markcite{taylor:09a}({Taylor} {et~al.} 2009a). \\markcite{wuyts:07}{Wuyts} {et~al.} (2007) (hereafter W07) and \\markcite{williams:09}{Williams} {et~al.} (2009) (hereafter W09) demonstrate that quiescent galaxies up to $z\\sim2$ can be fairly cleanly separated from the actively star-forming population when multiple rest-frame colors are used. In this Letter we use the NEWFIRM Medium Band Survey \\markcite{nmbs}(NMBS; {van Dokkum} {et~al.} 2009) to explore the evolution of the galaxy color bimodality over $01.4$ at $z>1.2$. Conversely, there are 2826 objects that satisfy these criteria in the full photometric sample. The results presented here are consistent with those of W09, who show that quiescent galaxies can be separated from dusty star-forming galaxies in the $U~-~V$ vs. $V~-~J$ color-color diagram out to $z=2$. The reddening-corrected dead sequence galaxies discussed here fall nicely within the $UVJ$ quiescent selection criteria, though the reddening correction is critical for detecting the bimodal color distribution in a color-mass diagram at $z>1.5$, explaining why it is not seen by W09. We find the color evolution of dead sequence galaxies to be consistent with that found by B04, who show that it can be roughly explained by simple stellar populations formed at $z>2$ that age passively to the present day. We show, however, that additional care must be taken when defining red sequence galaxy samples as done by, e.g., B04. The bottom panel of Figure \\ref{f:average_av} shows that a red galaxy selection based on a single rest-frame color will contain a significant fraction of dusty-starburst galaxies ($\\sim$20\\% with $A_V>1$) that would otherwise be much bluer (see also \\markcite{cimatti:02}{Cimatti} {et~al.} 2002; \\markcite{brammer:07}{Brammer} \\& {van Dokkum} 2007; W07; \\markcite{gallazzi:09}{Gallazzi} {et~al.} 2009; W09; \\markcite{wolf:09}{Wolf} {et~al.} 2009). The reddening correction dramatically reduces the number of objects in the green valley, so such a correction would be necessary when using those objects to estimate the rate at which galaxies move from the blue to red sequences \\markcite{martin:07}({Martin} {et~al.} 2007). For example, a sparsely-populated green valley at most redshifts could constrain the rate at which star-forming galaxies are quenched, which in turn may help constrain the processes that regulate star formation in theoretical models. As might be expected, the large majority of objects detected at \\24mum\\ have reddening-corrected colors that place them on the blue-sequence \\markcite{reddy:06}({Reddy} {et~al.} 2006, see also Figure 3). A number of MIPS-detected galaxies, however, remain on the red sequence even after applying the reddening correction: roughly one in seven MIPS objects have dust-corrected colors within $0.5$ mag of the red sequence ridgeline drawn in Figure \\ref{f:uv_z_av}. The dead sequence galaxies detected with MIPS have de-reddened colors that are slightly bluer than those without MIPS detections (Figure \\ref{f:cmd}). This could perhaps be due to low levels of recent star formation or to AGN contamination of the UV-optical light, though this tentative result requires more detailed investigation because, for example, it is not clear that the dust correction or even the extinction law itself are appropriate for these galaxies. Figures \\ref{f:uv_z_av} and \\ref{f:cmd} demonstrate that the dead sequence persists to $z=2\\mbox{--}2.5$, which is consistent with the discovery of a small (spectroscopic) sample of red sequence objects with low sSFR at $z\\sim2.3$ described by \\markcite{kriek:08}{Kriek} {et~al.} (2008). Such studies of red sequence galaxies found among the field are complementary to others that target the higher density environments of clusters at $z\\gtrsim1$ \\markcite{mei:09}(e.g., {Mei} {et~al.} 2009) and proto-clusters at $z\\sim2$ \\markcite{zirm:08}({Zirm} {et~al.} 2008). Mechanisms that have been proposed to suppress star formation in massive galaxies, including the much-discussed feedback from AGN and gravitational heating from cosmological accretion \\markcite{naab:07,dekel:08}({Naab} {et~al.} 2007; {Dekel} \\& {Birnboim} 2008), must be able to explain the formation of the dead sequence in a variety of environments as early as $z\\sim2.5$. Finally, we note that quantitative constraints on the assembly and evolution of passively-evolving galaxies require studies of the evolution of their mass function, possibly combined with structural information (see, e.g., B04; \\markcite{faber:07}{Faber} {et~al.} 2007; \\markcite{cimatti:08}{Cimatti} {et~al.} 2008; \\markcite{vandokkum:08}{van Dokkum} {et~al.} 2008, and many other studies). The NMBS is ideally suited to help disentangle these processes by providing accurate photometric redshifts and colors over a large area in the redshift regime of the most rapid build-up of massive galaxies." }, "0910/0910.0014_arXiv.txt": { "abstract": "Finding electromagnetic (EM) counterparts of future gravitational wave (GW) sources would bring rich scientific benefits. A promising possibility, in the case of the coalescence of a super-massive black hole binary (SMBHB), is that prompt emission from merger-induced disturbances in a supersonic circumbinary disk may be detectable. We follow the post--merger evolution of a thin, zero-viscosity circumbinary gas disk with two-dimensional simulations, using the hydrodynamic code FLASH. We analyze perturbations arising from the 530 ${\\rm km~s^{-1}}$ recoil of a $10^6 M_\\odot$ binary, oriented in the plane of the disk, assuming either an adiabatic or a pseudo--isothermal equation of state for the gas. We find that a single-armed spiral shock wave forms and propagates outward, sweeping up $\\sim 20\\%$ of the mass of the disk. The morphology and evolution of the perturbations agrees well with those of caustics predicted to occur in a collisionless disk. Assuming that the disk radiates nearly instantaneously to maintain a constant temperature, we estimate the amount of dissipation and corresponding post-merger light--curve. The luminosity rises steadily on the time--scale of months, and reaches few $\\times 10^{43}$ erg/s, corresponding to $\\approx 10\\%$ of the Eddington luminosity of the central SMBHB. We also analyze the case in which gravitational wave emission results in a $5\\%$ mass loss in the merger remnant. The mass-loss reduces the shock overdensities and the overall luminosity of the disk by $\\approx 15-20\\%$, without any other major effects on the spiral shock pattern. ", "introduction": "The gravitational waves (GWs) produced during the late stages of the merger between super-massive black holes (SMBHs), with masses of $\\sim (10^4$--$10^7)\\,{\\rm M_\\odot}/(1+z)$, out to redshifts beyond $z\\approx 10$, are expected to be detectable in the next decade by the {\\it Laser Interferometric Space Antenna} ({\\it LISA}) satellite. While the GW signatures themselves will be a rich source of information, identifying the electromagnetic (EM) counterpart of the LISA source would open up a whole new range of scientific opportunities, from black hole astrophysics to fundamental aspects of gravitational physics and cosmology \\citep{koc07,HKM09a,Phinney,Bloom}. Whether the EM counterparts of the SMBHBs detected by {\\it LISA} can be uniquely identified depends on the accuracy of localization by {\\it LISA}, and on the nature of the EM emission. For the typical SMBHB detected by {\\it LISA} at $z\\approx 1-2$, the sky position and the redshift will be known to the accuracy of $\\delta(\\Delta\\Omega)\\sim$ few $\\times$ 0.1 square degrees, and $\\delta z\\approx$ few $\\times$ 0.01, respectively (the latter limited by weak lensing errors; e.g. \\citealt{hh05,koc06}). Within this three--dimensional error volume, there will be of order $\\sim 100$ candidate galaxies, at the optical magnitude limit expected to host such SMBHBs \\citep{koc08}. If the coalescing SMBHs themselves produce bright emission comparable to luminous quasars, or are associated with some other, similarly rare subset comprising $\\lsim 1\\%$ of all galaxies (such as ultra--luminous infrared galaxies), then a unique counterpart may be identified among these candidates \\citep{koc06}. However, having a prediction for the spectrum and the light--curve of a coalescing SMBHB -- and therefore knowing what characteristic signatures to look for -- would make such identifications both more likely and more reliable. A promising possibility is that the coalescing SMBHB produces a {\\em variable or transient} EM signal, which can be uncovered by suitably designed EM observations, either concurrently with or following the {\\it LISA} detection \\citep{koc08,lh08,HKM09a}. The dense nuclear gas around the BH binary is expected to cool rapidly, and settle into a rotationally supported, thin circumbinary disk \\citep[e.g.][]{barnes,elcm05}. The evolution of a SMBHB embedded in such a disk has been studied in various idealized configurations \\citep[e.g.][]{an02,liu03,mp05,dot06,mm08,hayasaki09,cuadra09,HKM09b}. Generically, at small orbital separations when the binary is detectable by {\\it LISA}, the orbit decays rapidly due to GW emission. If the disk is thin, the binary torques create a central cavity, nearly devoid of gas, within a region about twice the orbital separation \\citep[e.g.][]{al94}. Whether the decaying SMBHBs produce bright emission during this stage is not well understood. If the central cavity were truly empty, no gas would reach the SMBHBs, and any emission produced farther out in the disk would likely be weak. On the other hand, numerical simulations suggest residual gas inflow into the cavity \\citep{al96,mm08,hayasaki07,hayasaki08,cuadra09}, which may plausibly accrete onto the BHs, producing non--negligible EM emission. A different possibility, and the focus of the present paper, is that variable EM signatures are produced in the gas disk {\\it promptly after} the coalescence of the SMBHB. The burst of GWs emitted during the last stages of the coalescence results in a corresponding, nearly instantaneous reduction in the binary's rest mass. Furthermore, when compact objects coalesce asymmetrically, the linear momentum carried away by the GWs imparts a ``kick'' to the center of mass of the system \\citep{Bekenstein, Fitchett}. Recent break--through in numerical relativity has allowed accurate calculations of these effects, showing that the mass loss is typically several percent \\citep[e.g.][and references therein]{tm08}, while kick velocities are typically several hundred ${\\rm km~s^{-1}}$, but can be as high as 4,000 ${\\rm km~s^{-1}}$ \\citep{baker06,baker07,baker08,campanelli07a,campanelli07b,gonzalez07a,gonzalez07b, herr07a,herr07b,herr07c,koppitz07}. The circumbinary gas will respond promptly (on the local orbital timescale) to such dynamical disturbances. Importantly, the gaseous disk outside the inner cavity is expected to cool efficiently and become geometrically thin, implying that the orbital motion of the gas is supersonic. This gas is therefore susceptible to prompt shocks, which could, in principle, produce a detectable transient EM signature \\citep{mp05}. \\citet{Lippai} considered the motion of collisionless test particles in such a disturbed disk. As long as the particles remain bound to the central SMBHB, they follow elliptical Kepler orbits (in the inertial frame of the SMBHB). These orbits cross, and produce a characteristic, outward--propagating spiral caustic. While these results were obtained in a pressureless ``dark matter'' disk, they suggest that shocks, with a similar pattern, will indeed arise in the gas, on times--scales of $\\sim$weeks to $\\sim$months after the coalescence. Similar conclusions were reached by \\citet{Shields} and by \\citet{Schnittman}, using N-body simulations and semi--analytical arguments, respectively, to show that a bright, post--merger ``flare'' may occur. Both of these studies focused on the evolution of disks around more massive BHs on longer ($\\sim 10^4$yr) time--scales, and proposed detecting the flare by monitoring a population of active galactic nuclei (AGN). By comparison, our analysis here focuses on the disks around lower--mass BHs and on the shorter time--scales of several weeks to a year, which are relevant for follow--up observations triggered by {\\it LISA} detections. Motivated by the above, in the present paper, we follow up on earlier work, and compute the response of the gas disk including the effects of gas pressure using hydrodynamical simulations. Our main goals are (i) to assess shock formation in hydrodynamical disks, and (ii) to estimate the amount of dissipation and the resulting light--curve produced by the disk. In particular, for the latter, we will use simulations with a pseudo--isothermal equation of state. This corresponds implicitly to strong dissipation, and should represent an approximate upper limit on the disk luminosity. Our computations are performed with the publicly available code FLASH \\citep{Fryxell}. Two other, independent studies have recently used 3--dimensional simulations to address the response of gas disks to mass--loss \\citep{ONeill} and to both mass--loss and kicks \\citep{Megevand}. The main difference between our analysis and these previous works is our inclusion of runs with a pseudo--isothermal equation of state, which modifies, qualitatively, the expected EM signature. In particular, \\citet{ONeill} and \\citet{Megevand} both run simulations with an adiabatic equation of state, and find that in most cases, the gas disk typically {\\em dims} following the merger. We observe a similar effect in our adiabatic runs, and attribute it to an overall dilution of the gas density. However, the gas develops significantly larger density contrasts in our pseudo--isothermal runs, and we argue that this can result in a significant {\\it brightening} of the system. Another important difference between our study and those of \\citet{ONeill} and \\citet{Megevand} is that we perform two--dimensional simulations, whereas \\citet{ONeill} and \\citet{Megevand} both utilized 3D simulations. While we have sacrificed resolving the vertical disk structure, this allows us to simulate a disk that extends to much larger radii ($10^4$ Schwarzschild radii), and follow the entire disk for a much longer period ($\\sim1$ yr for a $10^6~{\\rm M_\\odot}$ binary), in order to cover the regime relevant for follow-up observations of {\\it LISA} events. Given the expected size of the inner cavity in the disk, $\\sim 100 R_s$, studying these outer regions of the disk, and following the disk evolution on the correspondingly longer time--scale, is especially important. Further differences between our study and those of \\citet{ONeill} and \\citet{Megevand} will be discussed in \\S~\\ref{sec:Others} below. The rest of this paper is organized as follows. In \\S~\\ref{sec:ProblemSetup}, we describe our computational setup, including details of the simulations, and initial conditions. In \\S~\\ref{sec:Results}, we present and discuss our main results. These include the impact of the kick (\\S~\\ref{sec:PostMerger}) and the mass--loss (as well as both effects in combination; \\S~\\ref{sec:MassLoss}), using constant surface density disks. In \\S~\\ref{sec:Observational}, we discuss our estimate of the post--merger light curves. In \\S~\\ref{sec:Alphadisks}, we repeat our calculations at higher resolution and with more realistic initial gas density and temperature profiles (adapted from a standard $\\alpha$-disk). In \\S~\\ref{sec:Numerical}, we discuss possible numerical issues, and in \\S~\\ref{sec:Others}, we compare our results to previous work. Finally, in \\S~\\ref{sec:Conclude}, we summarize our main conclusions and their implications, and outline natural future extensions of this study. ", "conclusions": "\\label{sec:Conclude} In this paper, we used two--dimensional hydrodynamical simulations of a circumbinary disk, to follow the effects of a velocity recoil and mass-loss of the central black hole binary, following the merger of the two black holes. From a suite of runs, we are able to draw two basic conclusions. First, the outward--propagating spiral shocks that develop in our simulations follow a pattern very similar to the caustics identified in collisionless disk \\citep{Lippai}. Gas pressure has a modest overall impact on the propagation and on the overall morphology of the shocks. On the other hand, we find that the gas pressure, and the assumed equation of state, has a strong impact on the overdensities that develop and on the strengths of the shocks: isothermal, low--temperature disks have larger overdensities and stronger shocks. Second, we have estimated an upper limit on the luminosity emerging from the disk experiencing a BH kick with $v_{\\rm kick}=530~{\\rm km~s^{-1}}$, by measuring the effective dissipation that occurs, implicitly, in the simulations when an isothermal condition is imposed on the gas. The resulting luminosity is of order $10\\%$ of the Eddington limit for the $10^6 M_\\odot$ BH used in the simulation, which suggests that the after--glow may be bright enough to be detectable. If the pre-merger disk has a luminosity below that of a standard steady-state thin accretion disk (due to the evacuation of the inner regions of the disk), then the merger-induced kick will cause a significant (order-of-magnitude, or larger) brightening of the disk. We also estimated the effective black--body temperature of the radiation emerging from the optically thick disk. We found that as the disk brightens, the characteristic frequency increases with time, possibly offering a unique signature of the kick--induced emission. When the accompanying mass--loss of the merger remnant is included in our simulations, the density contrasts and luminosity from the spiral shocks decrease somewhat, but do not change dramatically in overall behavior. While our results are encouraging, and suggest that the EM signature of the disturbed circumbinary disk may be detectable, this conclusion has to be verified in the future in improved models. In particular, a better estimate of the thermodynamics of the disk, with a realistic treatment of radiative cooling, should reveal how close the luminosity is to the values we obtained here, using the isothermal assumption; by using three--dimensional simulations that resolve the vertical disk structure, and by following the vertical transfer of the radiation produced should clarify the robustness of our conclusions about the evolution of the characteristic emergent frequency." }, "0910/0910.0687_arXiv.txt": { "abstract": "Emission of high energy (HE) photons above 100 MeV that is delayed and lasts much longer than the prompt MeV emission has been detected from several long duration gamma ray bursts (LGRBs) and short hard bursts (SHBs) by the Compton, Fermi and AGILE gamma ray observatories. In this paper we show that the main observed properties of this HE emission are those predicted by the cannonball (CB) model of GRBs: In the CB model all the observed radiations in a GRB are produced by the interaction of a highly relativistic jet of plasmoids (CBs) with the environment. The prompt X-ray and MeV $\\gamma$-ray pulses are produced by inverse Compton scattering (ICS) of glory photons -photons scattered/emitted into a cavity created by the wind/ejecta blown from the progenitor star or a companion star long before the GRB- by the thermal electrons in the CBs. A simultaneous optical and high energy emission begins shortly after each MeV pulse when the CB collides with the wind/ejecta, and continues during the deceleration of the CB in the interstellar medium. The optical emission is dominated by synchrotron radiation (SR) from the swept-in and knocked-on electrons which are Fermi accelerated to high energies by the turbulent magnetic fields in the CBs, while ICS of these SR photons dominates the emission of HE photons. The lightcurves of the optical and HE emissions have approximately the same temporal behaviour but have different power-law spectra. The emission of very high energy (VHE) photons above 100 TeV is dominated by the decay of $\\pi^0$'s produced in hadronic collisions of Fermi accelerated protons in the CBs. The CB model explains well all the observed radiations, including the high energy radiation from both LGRBs and SHBs as demonstrated here for GRB 090902B and SHB 090510. ", "introduction": "During nearly 20 years after the launch of the Compton Gamma Ray Observatory (CGRO), the Burst And Transient Source Experiment (BATSE) on board the CGRO has detected and measured light curves and spectra (Kaneko et al.~2008) in the sub-MeV range of several thousands gamma ray bursts (GRBs). Higher-energy observations with the EGRET instrument aboard CGRO were limited to those GRBs which happened to be in its narrower field of view. Its large calorimeter measured the light-curves and spectra of several GRBs in the 1-200 MeV energy range. Seven GRBs were detected also with the EGRET spark chamber, sensitive in the 30 MeV - 10 GeV energy range. The EGRET detections indicated that the spectrum of bright GRBs extends beyond 1 GeV (Hurley et al.~1994) with no evidence for a spectral cut-off (see, e.g., Dingus~1995,~2001 and references therein). However, a few GRBs, such as 940217 (Hurley et al.~1994) and 941017 (Gonzalez et al.~2003), showed evidence that the high energy component has a slower temporal decay than that of the sub-MeV emission, suggesting that, at least in some cases, it is not a simple extension of the main component, but originates from a different emission mechanism and/or region. This has been confirmed recently by observations of high energy photons in the 30 MeV - 300 GeV range from several GRBs with the AGILE (GRB 080514B: Giuliani et al.~2008, GRB 090510: Giuliani et al.~2009) and the Fermi large area telescope (LAT) (e.g., GRB 080916C: Abdo et al.~2009a, GRB 090902B: Bissaldi et al.~2009 and GRB 090510: Ghirlanda et al.~2009). The arrival times of the high energy photons did not coincide with the times of the brightest peaks seen at hard X-rays and MeV $\\gamma$-rays. Also the high energy emission lasted much longer time than that of the prompt keV-MeV emission. The detection of higher energy gamma rays is affected by pair production in their collisions with the extragalactic infrared background light (Nikishov 1961) resulting in an absorption which is a strong function of redshift and energy. Recent estimates (Primack et al.~2005), validated by HESS observations (Aharonian et al.~2006) predict an optical depth of roughly unity to 500 GeV photons emitted at a redshift $z\\!=\\!0.2$ and to 10 TeV photons at $z=0.05$. The average redshifts of LGRBs and SHBs are much larger, $z\\!=\\!2.2$ and $z\\!=\\!0.5$, respectively. Despite of the strong attenuation of high energy gamma rays in the intergalactic medium (IGM), there have been several claims in the past of detections at the 3 sigma level of TeV gamma-rays from GRBs (see, e.g., Atkins et al.~2003 and references therein). However, more recently, no GRB was conclusively detected in the range 100 GeV to 100 TeV by the ground based water Cherenkov detector Milagro and by the air Cherenkov telescopes MAGIC, Whipple, HESS and VERITAS. Moreover in all previous cases of reported detection of TeV gamma-rays from a GRB, the GRB redshift was not known. If their redshifts are similar to those of ordinary GRBs then their TeV gamma-rays are strongly absorbed by the extragalactic background light (EBL), implying that the TeV emission if detected must be extraordinarily energetic, i.e., with a much larger fluence than that emitted in X-rays and MeV $\\gamma$-rays. Most theoretical models of high energy photon emission in GRBs relied on the standard fireball models of GRBs (for recent reviews see, e.g., Piran~2005; M\\'{e}sz\\'{a}ros~2006; Zhang~2007). In these models, synchrotron radiation of electrons accelerated by `internal shocks' in collisions between conical shells ejected by the GRB's central engine produces the prompt GRB emission and a blast wave (external shock) driven into the circumburst medium generates their afterglow. The high energy radiation was suggested to be produced either by inverse Compton scattering of the synchrotron radiation (the so called `synchrotron self Compton mechanism') by the shock-accelerated electrons (e.g., Dermer et al.~2000), by the decay of $\\pi^0$ photo produced in collisions of shock accelerated hadrons with synchrotron photons in the expanding fireball (Waxman \\& Bahcall~1997), or by synchrotron radiation from ultra high energy protons (Totani~1998) allegedly accelerated in the GRB fireball (Waxman~1995; Milgrom \\& Usov~1995; Vietri~1995). All these models that were based on the standard fireball model of GRBs predict simultaneous emissions at all energy bands. This is in conflict with the observed delayed emission of the high energy photons that lasts much longer. In the cannonball (CB) model (see e.g. Dado, Dar \\& De R\\'ujula, hereafter DDD, 2009a,b and references therein), GRBs and their afterglows (AGs) are produced by bipolar jets of highly relativistic plasmoids (CBs) ejected in violent stellar processes. The prompt MeV $\\gamma$-rays and hard X-rays are produced by inverse Compton scattering (ICS) of glory photons - photons emitted/scattered into a cavity formed by the wind/ejecta puffed by the progenitor or a companion star long before the GRB. Synchrotron radiation (SR) emitted from the electrons of the ionized wind/ejecta that are swept into the CBs and are Fermi accelerated by their turbulent magnetic fields dominates their `prompt' optical emission (e.g., DDD2009a and references therein) which begins when the CBs reach the wind/ejecta. In this paper we show that in the CB model ICS of this SR (e.g., Dado \\& Dar~2005) dominates the HE emission from long GRBs and SHBs. This HE emission begins simultaneously with the `prompt' optical emission that lags after the prompt X-ray and MeV $\\gamma$-ray emission and lasts much longer, has the same lightcurve as the optical emission but with a power-law spectrum that is identical to that of the X-ray emission. It describes well the HE observations, as demonstrated in the paper for GRB 090902B and SHB 090510. Because of the Klein-Nishina effect, $E^2\\,dn/dE$ of ICS of the prompt SR in GRBs peaks near 10 TeV and is cutoff at a few tens of TeV when the rate of synchrotron energy losses by electrons in the CBs exceeds the rate of energy gain by Fermi acceleration. The prompt decay of $\\pi^0$'s produced in the collisions between the Fermi accelerated nuclei and the ambient matter in the CBs produces a power-law spectrum that extends to much higher energies where it dominates the HE emission. However, like in blazars, the observed flux of TeV photons from distant GRBs and SHBs is stronglyly suppressed by pair production in collisions with the extragalactic background photons and only relatively nearby GRBs and SHBs might be detected in TeV photons by the large ground based HE gamma ray telescopes such as HESS, MAGIC and VERITAS. ", "conclusions": "In the cannonball (CB) model, high energy emission from GRBs and SHBs is a natural consequence of the model. The dominant leptonic and hadronic emission mechanisms are ICS of SR by Fermi accelerated electrons in the collision of the jet of highly relativistic CBs with the wind/ejecta blown from the progenitor or companion star long before the GRB, and the decay of $\\pi^0$ produced in hadronic collisions between the CB nuclei and the nuclei of the hadronic matter (ejecta/wind/ISM) that the jet passes through. The main predictions of the model for the early-time high energy emission are: \\begin{itemize} \\item{} Each prompt keV-MeV pulse is followed by a delayed HE emission which lasts much longer. \\item{} The HE emission coincides in time with the optical emission. \\item{} The light curve of the HE emission is roughly proportional to that of the unextinct optical emission. The decay of both the `prompt' optical and the HE emissions is a power-law with an index $\\alpha$=1+$\\beta_0\\!\\sim 1.5$. \\item{} The spectrum of the HE component is a simple power-law with a spectral index approximately equal to that of the X-ray afterglow, i.e., $\\Gamma_{LAT}\\!\\approx\\!\\Gamma_{X}$. \\item{} The HE emission extends to very high energies, where the observed radiation is strongly attenuated by $e^+e^-$ pair productin in the EBL. \\item{} The equivalent isotropic energy of the HE component in bright GRBs can reach, and even exceed that in the sub-MeV range. \\item{} The neutrino counterpart of the hadronic emission of HE $\\gamma$-rays from the most luminous GRBs is barely detectable in $km^3$ underwater/under-ice Cherenkov neutrino telescopes (DD2008). \\end{itemize} \\noindent These predictions are consistent with the present available data on high energy emission from GRBs obtained from the gamma ray satellites and from the large air, ground and underground Cherenkov telescopes. In particular, the main observed properties of the HE emission measured with the Compton, Fermi and AGILE gamma ray satellites are well reproduced by the CB model as demonstrated in this paper for GRB 090902B and SHB 090510. An observational proof of the ICS origin of the HE gamma ray emission from GRBs requires simultaneous detections of the prompt optical and HE emissions. It is highly desireable that the LAT team improves their automatic analysis, in order to deliver a GRB position within a few seconds after the LAT detection. Even a few tens of seconds will be very useful. This will provide a stringent test of models of HE emission from GRBs (and from other HE transient sources such as blazars, microquasars, pulsars, etc) and help pin down their production mechanism. Extremely optically-luminous GRBs, such as GRB 080319 where the prompt optical emission was resolved into individual flares (Racusin et al.~2008), may show that also the prompt high energy emission consists of HE flares which are associated with and follow promptly each individual keV-MeV pulse, as expected in the CB model. \\newpage" }, "0910/0910.0478_arXiv.txt": { "abstract": "We construct one-zone steady-state models of cosmic ray (CR) injection, cooling, and escape over the entire dynamic range of the FIR-radio correlation (FRC), from normal galaxies to starbursts, over the redshift interval $0 \\le z \\le 10$. Normal galaxies with low star-formation rates become radio-faint at high $z$, because Inverse Compton (IC) losses off the CMB cool CR electrons and positrons rapidly, suppressing their nonthermal radio emission. However, we find that this effect occurs at higher redshifts than previously expected, because escape, bremsstrahlung, ionization, and starlight IC losses act to counter this effect and preserve the radio luminosity of galaxies. The radio dimming of star-forming galaxies at high $z$ is not just a simple competition between magnetic field energy density and the CMB energy density; the CMB must also compete with every other loss process. We predict relations for the critical redshift when radio emission is significantly suppressed compared to the $z \\approx 0$ FRC as a function of star-formation rate per unit area. For example, a Milky Way-like spiral becomes radio-faint at $z \\approx 2$, while an M82-like starburst does not become radio-faint until $z \\approx 10 - 20$. We show that the ``buffering'' effect of non-synchrotron losses improves the detectability of star-forming galaxies in synchrotron radio emission with EVLA and SKA. Additionally, we provide a quantitative explanation for the relative radio brightness of some high-$z$ submillimeter galaxies. We show that at fixed star formation rate surface density, galaxies with larger CR scale heights are radio bright with respect to the FRC, because of weaker bremsstrahlung and ionization losses compared to compact starbursts. We predict that these ``puffy starbursts'' should have steeper radio spectra than compact galaxies with the same star-formation rate surface density. We find that radio bright submillimeter galaxies alone cannot explain the excess radio emission reported by ARCADE2, but they may significantly enhance the diffuse radio background with respect to a naive application of the $z \\approx 0$ FRC. ", "introduction": "The radio and FIR luminosities of star-forming galaxies are linearly correlated over three decades in luminosity \\citep[][]{vanDerKruit71,vanDerKruit73,deJong85,Helou85,Condon92,Yun01}. This FIR-radio correlation (FRC) is driven by star-formation. Starlight emitted by young, massive stars in the optical and UV is reprocessed by dust into far-infrared radiation. At the same time, the supernova remnants of massive stars accelerate cosmic rays (CRs), which produce synchrotron radio emission in the galaxies' magnetic fields. The relation at $z \\approx 0$ is linear except in the very lowest luminosity galaxies \\citep{Condon91b,Yun01,Bell03}. Since the FRC has less than a factor of 2 scatter in the local Universe, radio luminosity is often used as a tracer of star-formation at both low and high redshift \\citep[e.g.,][]{Cram98,Mobasher99}. Although the FRC has been mainly studied in the low-$z$ universe, there has been recent interest in understanding how it applies to star-forming galaxies at high $z$. There have been several conflicting results observationally. A number of studies have found that the FRC holds unchanged or with little evolution at high redshifts (e.g., \\citealt{Appleton04,Ibar08,Rieke09,Murphy09,Younger09,Garn09,Sargent10,Ivison10,Younger10}; \\citealt{Persic07} suggest radio-dim ULIRGs at high $z$). Submillimeter galaxies, however, seem to be radio-bright, by a factor of $\\sim 3$ (\\citealt{Kovacs06,Vlahakis07,Sajina08,Murphy09,Seymour09,Murphy09c,Michalowski09,Michalowski10}). In particular, \\citet{Murphy09} and \\citet{Michalowski09} argue that their samples of submillimeter galaxies are intrinsically radio bright with respect to the local FRC, with no significant contamination from radio bright AGN. In addition to the work on individual star-forming galaxies, there have also been new investigations into both the recently-resolved FIR and the unresolved GHz radio backgrounds. The FIR background was detected by COBE and has long been attributed to star-formation \\citep[see the reviews by][]{Hauser01,Lagache05}. The star-formation origin of the FIR background was recently confirmed by BLAST \\citep{Devlin09,Pascale09}. A detection of the extragalactic GHz radio background has also been recently reported by ARCADE2 \\citep{Fixsen09,Seiffert09}. Predictions for the strength of the diffuse radio background from star formation rely on the FRC holding out to $z \\approx 1 - 2$ \\citep[e.g.,][]{Haarsma98,Dwek02}, where most star-formation occurs. In this paper, we describe how and why the linearity and normalization of the FRC are broken (or preserved) for different types of galaxies at high $z$. We describe the basic physics of our model and the elements of the high-$z$ universe that physically allow the FRC to break down in \\S~\\ref{sec:Theory}. In \\S~\\ref{sec:Procedure}, we briefly describe our underlying procedure. In \\S~\\ref{sec:Results}, we discuss our results, providing predictions for when star-forming galaxies should be more or less radio luminous than the $z \\approx 0$ FRC suggests. We also summarize the implications of our predictions for radio emission as a star formation tracer in \\S~\\ref{sec:SFTracer} and estimate the effects of radio-bright SMGs on the diffuse radio background in \\S~\\ref{sec:RadioBackground}. Throughout this paper, primed quantities denote the source's rest-frame and unprimed quantities are for the observer-frame. ", "conclusions": "\\label{sec:Results} \\label{sec:Discussion} \\subsection{The Evolving FIR-Radio Correlation} \\label{sec:FRCEvolution} \\begin{figure*} \\centerline{\\includegraphics[width=9cm]{f1a.eps}\\includegraphics[width=9cm]{f1b.eps}} \\figcaption[simple]{The FIR-radio correlation at high redshift, in the rest-frame, using the B07 star formation law. On left is the case with $B \\propto \\Sigma_g^{0.7}$ and no winds; on right, the case with $B \\propto \\rho^{0.6}$ and winds. Solid lines have $h = 100~\\pc$ for starbursts ($\\Sigma_g \\ge 0.1~\\gcm2$; $\\Sigma_{\\rm SFR} \\ge 2 - 4~\\Msun~\\kpc^{-2}~\\yr^{-1}$), while dotted lines have $h = 1~\\kpc$ for starbursts. Both panels show that a linear FRC is produced at $z = 0$ with the correct normalization. Galaxies become radio-dim at high redshift as IC losses off the CMB increase. Note that in each case the puffy starbursts do not lie on the same FRC as compact starbursts: there is a systematic offset caused by the unbalancing of the high-$\\Sigma_g$ conspiracy (\\S~\\ref{sec:Puffy}). We do not include thermal radio emission, which will set a minimum radio luminosity or maximum $q^{\\prime}_{\\rm FIR}$ at each $\\Sigma_g$.\\label{fig:LFIRRadioRest}} \\end{figure*} We show the rest-frame FRC for our model with $B \\propto \\Sigma_g^{0.7}$, the B07 star-formation law, and no winds, in Figure~\\ref{fig:LFIRRadioRest} (\\emph{left panel}) as an example. The solid dark red line is the $z = 0$ FRC. All solid lines assume that starbursts are compact, with $h = 100~\\pc$. It is clear from Figure~\\ref{fig:LFIRRadioRest} that compact starbursts show little evolution in the FRC, while low surface density galaxies have lower radio luminosities at high redshift. This behavior is robust in all of the variants. The cause of evolution in the FRC is Inverse Compton losses off the CMB for CR electrons and positrons. Figure~\\ref{fig:FRCEvolution} (\\emph{left panel}) shows that the FRC should display relatively little evolution out to $z \\approx 1$, except for the lowest surface brightness galaxies. However, normal galaxies have suppressed synchrotron radio emission at $z \\approx 2$, a factor of $\\sim 2$ for $\\Sigma_g = 0.01~\\gcm2$ ($\\Sigma_{\\rm SFR} \\approx 0.06~\\Msun~\\kpc^{-2}~\\yr^{-1}$) and of order 10 for $\\Sigma_g = 0.001~\\gcm2$ ($\\Sigma_{\\rm SFR} \\approx 0.001 - 0.002~\\Msun~\\kpc^{-2}~\\yr^{-1}$). The radio luminosities continue to fall with redshift. At $z \\approx 5$, IC off the CMB starts to matter even for the weaker starbursts ($\\Sigma_g = 0.1~\\gcm2$; $\\Sigma_{\\rm SFR} \\approx 2 - 4~\\Msun~\\kpc^{-2}~\\yr^{-1}$). Dense starbursts ($\\Sigma_g = 10~\\gcm2$; $\\Sigma_{\\rm SFR} \\approx 900~\\Msun~\\kpc^{-2}~\\yr^{-1}$ for the K98 law or $9000~\\Msun~\\kpc^{-2}~\\yr^{-1}$ for the B07 law) remain on a linear FRC even at $z \\approx 10$. The strong cooling from bremsstrahlung, ionization, and IC off starlight implied by the high-$\\Sigma_g$ conspiracy (\\S~\\ref{sec:Theory}) acts as a buffer against IC losses off the CMB. Similarly, diffusive escape provides a similar buffering effect in low-$\\Sigma_g$ galaxies, so that the increased IC losses off the CMB does not suppress the radio emission quickly. In other words, most of the power from IC scattering of CMB photons does not come at the expense of synchrotron, but of other losses. The radio dimming cannot be described by a simple competition between magnetic field energy density and the CMB energy density; the CMB energy density must also compete with \\emph{every other loss process}. The buffering actually serves as an important test for both conspiracies described in \\S~\\ref{sec:Theory}. In high-$\\Sigma_g$ galaxies and starbursts, we predict that bremsstrahlung, ionization, and Inverse Compton of starlight already take a large portion of a \\GHz~electron's energy budget; hence, the high-$\\Sigma_g$ conspiracy works to hold $L_{\\rm TIR}^{\\prime}/L_{\\rm radio}^{\\prime}$ constant out to quite high redshift. The buffering essentially doubles the redshift that compact starbursts remain on the FRC; the weakest starbursts ($\\Sigma_g = 0.1~\\gcm2$; $\\Sigma_{\\rm SFR} \\approx 2 - 4~\\Msun~\\kpc^{-2}~\\yr^{-1}$) are radio-dim by a factor of $3$ at $z = 10$ with buffering, instead of $z = 5$ without the buffering. In low-$\\Sigma_g$ galaxies, electrons and positrons can easily escape before they lose energy to Inverse Compton off the CMB at low enough $z$; thus, the low-$\\Sigma_g$ conspiracy also works to hold $L_{\\rm TIR}^{\\prime}/L_{\\rm radio}^{\\prime}$ constant. The suppression due to IC losses off the CMB seen in Figure~\\ref{fig:LFIRRadioRest} can be estimated by examining the ratios of the synchrotron cooling time to the total loss time, including both escape and cooling losses. In Appendix~\\ref{sec:CMBRadioDim}, we derive the ratios of loss times and we show that for a given Schmidt law, a critical redshift $z_{\\rm crit}$ can be defined for each $\\Sigma_g$ and $\\Sigma_{\\rm SFR}$ at which radio emission is suppressed. We define $z_{\\rm crit}$ to be the redshift when the rest-frame radio luminosity at $\\nu^{\\prime} = 1.4~\\GHz$ is suppressed by a factor of 3 compared to $z = 0$. For our model with the B07 star-formation law, no winds, and $B \\propto \\Sigma_g^{0.7}$, we find the critical redshifts are approximated as \\begin{equation} \\label{eqn:B07zCrit} z_{\\rm crit} \\approx \\left\\{ \\begin{array}{ll} 1.4 & ({\\rm Normal~galaxies}, \\Sigma_{\\rm SFR} \\la 0.02)\\\\ 5.8 \\Sigma_{\\rm SFR}^{0.23} - 1 & ({\\rm Normal~galaxies}, \\Sigma_{\\rm SFR} \\ga 0.02)\\\\ 5.7 \\Sigma_{\\rm SFR}^{0.23} - 1 & ({\\rm Puffy~starbursts})\\\\ 7.4 \\Sigma_{\\rm SFR}^{0.23} - 1 & ({\\rm Compact~starbursts}), \\end{array} \\right. \\end{equation} where $\\Sigma_{\\rm SFR}$ is in units of $\\Msun~\\kpc^{-2}~\\yr^{-1}$. We refer the reader to Appendix~\\ref{sec:CMBRadioDim}, where we present similar relations for our other models of the FRC. We note that the radio suppression could, conceivably, be used to measure the temperature of the CMB at high redshift. In principle, this method could apply to any galaxy with a radio and FIR detection. However, the conspiracies would have to be accounted for, and they are affected significantly by both the gas surface density $\\Sigma_g$ and scale height $h$ (see \\S~\\ref{sec:Puffy}). Any measurement of the CMB temperature would depend on assumptions of the galaxy properties and would be model-dependent. \\begin{figure*} \\centerline{\\includegraphics[width=9cm]{f2a.eps}\\includegraphics[width=9cm]{f2b.eps}} \\figcaption[simple]{The evolution of the FIR-radio correlation, both in the rest-frame at $\\nu^{\\prime} = 1.4~\\GHz$ (\\emph{left panel}), and as inferred for the rest-frame from observations at $\\nu = 1.4~\\GHz$ (\\emph{right panel}). Black lines have $h = 100~\\pc$ for starbursts, while grey lines have $h = 1~\\kpc$ for starbursts. Both panels show the case with $B \\propto \\Sigma_g^{0.7}$, no winds, and the B07 star-formation law. Normal galaxies become radio dim because of IC losses off the CMB at intermediate redshift. Both compact and puffy starbursts maintain their rest-frame radio luminosities until high redshift. All galaxies are ``buffered'' by non-synchrotron losses (\\S~\\ref{sec:FRCEvolution} and Appendix~\\ref{sec:CMBRadioDim}), so that the evolution is not as great as would be expected with only synchrotron and IC losses off the CMB. The inferred evolution is usually greater than the rest-frame evolution, because normal galaxies and puffy starbursts have $\\alpha \\ga 0.7$. \\label{fig:FRCEvolution}} \\end{figure*} As Figure~\\ref{fig:FRCEvolution} (\\emph{right panel}) shows, the inferred rest-frame values of $L^{\\prime}_{\\rm TIR}/L^{\\prime}_{\\rm radio}$ show additional apparent evolution, simply because the spectral slopes of the galaxies are not exactly $0.7$. Compact starbursts have flatter spectra\\footnote{In this paper (unlike LTQ), $\\alpha^{\\prime} \\equiv \\frac{d \\log F^{\\prime}_{\\nu}}{d \\log \\nu^{\\prime}}$ and $\\alpha \\equiv \\frac{d \\log F_{\\nu}}{d \\log \\nu}$ are instantaneous spectral slopes, not the measured spectral slopes between two observed frequencies, unless otherwise noted. \\label{ftnt:AlphaNotation}} with $\\alpha \\approx 0.4 - 0.6$ at 1.4 GHz, so by adopting $\\alpha = 0.7$ they \\emph{appear} to become slightly radio brighter until $z \\approx 1$, after which at higher frequencies their spectra steepen, and their radio emission appears to dim again at higher $z$. Normal galaxies have steeper spectra with $\\alpha \\approx 0.9 - 1.0$ at 1.4 GHz, so their apparent radio luminosity sinks below their true radio luminosity at high redshift. At higher redshifts, our models predict that galaxies have intrinsically steeper radio spectra, because of increased IC losses from the CMB. The rest-frame spectral slope $\\alpha^{\\prime}$ of normal galaxies at 1.4 GHz and asymptotes at $\\sim 1.1$ by $z \\approx 2 - 3$, when losses are dominated by IC for our injection spectrum of CRs with $p = 2.2$. There is much less intrinsic rest-frame evolution of starburst radio spectra; even at $z \\approx 5$, $\\alpha^{\\prime}$ increases only by 0.1 for the weakest compact starbursts. The observable $\\alpha^{1.4~\\GHz}_{610~\\MHz}$ shows much more pronounced evolution with redshift for starbursts, increasing by $0.1$ to $z \\approx 1 - 2$, and $0.2$ to $z \\approx 4 - 5 $. This effect arises simply because we predict the CR electron and positron spectra steepen with rest-frame frequency as synchrotron and IC losses become stronger \\citep[e.g.,][LTQ]{Thompson06}: at higher redshift and fixed observing frequency, we are seeing higher energy electrons and positrons. An important effect that we do not consider is the thermal free-free radio emission. This will set a minimum total observed radio emission that is directly proportional to the star-formation luminosity. Thus, the true total radio deficit at GHz will not be as big as the synchrotron-only deficits plotted in Figures~\\ref{fig:LFIRRadioRest} and \\ref{fig:FRCEvolution}. Thermal free-free emission will also flatten the spectrum, especially at high frequencies. However, free-free emission is much fainter than the synchrotron luminosity at GHz frequencies, except in the faintest star-forming galaxies \\citep{Condon92,Hughes06}. \\subsection{The Radio Excess (or Deficit) of Puffy Starbursts: Submillimeter Galaxies} \\label{sec:Submillimeter} \\label{sec:Puffy} As shown in Figure~\\ref{fig:LFIRRadioRest}, puffy starbursts fall on a linear FIR-radio correlation of their own (dotted lines), in line with observations of SMGs \\citep{Kovacs06,Murphy09}. The radio luminosity of these galaxies is nonetheless the result of a conspiracy, between IC losses on starlight, which decrease the radio luminosity, and the enhanced radio emission from secondary electrons and positrons, and the $\\nu_C$ effect (\\S~\\ref{sec:Theory}). The variation in $L_{\\rm TIR}^{\\prime}/L_{\\rm radio}^{\\prime}$ for puffy starbursts alone is usually less than a factor of $2$ over the range $0.1~\\gcm2 \\le \\Sigma_g \\le 10~\\gcm2$ (see Table~\\ref{table:Models}; in most variants the variation is $\\sim 1.6$). Like compact starbursts, escape plays essentially no role in most of the models, except that winds can slightly decrease the radio emission in relatively tenuous $\\Sigma_g = 0.1~\\gcm2$ ($\\Sigma_{\\rm SFR} \\approx 2 - 4~\\Msun~\\kpc^{-2}~\\yr^{-1}$) starbursts (footnote~\\ref{ftnt:WindVariant}). \\subsubsection{$B \\propto \\Sigma_g^{0.7-0.8}$: Radio-Bright Puffy Starbursts} It is plain from Figure~\\ref{fig:LFIRRadioRest} (\\emph{left panel}) that the normalization of the puffy starburst FRC (dotted lines) is different than the FRC of the compact starbursts and normal galaxies (solid lines) when $B \\propto \\Sigma_g^{0.7}$. We show in Table~\\ref{table:Models} that models with $B \\propto \\Sigma_g^{0.7-0.8}$ have radio-bright puffy starbursts compared to the observed local FRC, by a factor of $2 - 4$ at $z = 0$. Like the compact starbursts, puffy starbursts show little rest-frame evolution in $L_{\\rm TIR}^{\\prime}/L_{\\rm radio}^{\\prime}$, except at relatively low surface densities ($\\Sigma_g = 0.1~\\gcm2$; $\\Sigma_{\\rm SFR} \\approx 2 - 4~\\Msun~\\kpc^{-2}~\\yr^{-1}$), where they become radio dim at high redshifts because of IC losses on CMB photons. Therefore, we predict that puffy starbursts, which are mainly observed at high $z$, have \\emph{intrinsically} different radio properties not caused by their redshift. We propose a natural explanation of this radio excess in the framework of LTQ (\\S~\\ref{sec:Theory}). In dense starbursts, protons are efficiently converted into secondary electrons and positrons through inelastic proton-proton scattering, which contribute to the synchrotron emission. Furthermore, the $\\nu_C$ effect increases $L_{\\rm TIR}^{\\prime}/L_{\\rm radio}^{\\prime}$ for starbursts which have larger $B$. Compact starbursts lying on the $z = 0$ FRC balance these effects with increased bremsstrahlung, ionization, and IC losses, which compete with the synchrotron losses and suppress the excess radio luminosity (\\S~\\ref{sec:Theory}). In puffy starbursts with relatively low volume densities compared to their compact cousins at fixed $\\Sigma_g$, however, bremsstrahlung and ionization are not strong enough to compensate for these effects. Only synchrotron and IC losses remain, upsetting the conspiracy. The radio excess predicted by this picture is systematically greater when using the K98 star-formation relation, because the IC loss rate on starlight is smaller at fixed surface density by a factor of 2.5 to 11 from weak starbursts to the densest starbursts. This freedom to vary the radio-excess with $B$ does not exist in the standard calorimeter model, where all CR electron/positron energy goes into synchrotron emission, and the radio emission saturates. The balance between synchrotron and the other forms of cooling can be changed in starbursts to alter the normalization of the FRC, even though escape is negligible. There is a reason to expect our allowed $B \\propto \\Sigma_g^{0.7-0.8}$ specifically: radiation pressure may drive turbulence and enhance the magnetic field until its energy density is comparable to radiation \\citep[e.g.,][]{Thompson08}. If the K98 relation holds, then since the magnetic energy density scales as $U_B \\propto U_{\\rm ph}$, $B \\propto \\Sigma_g^{0.7}$, and if the B08 relation holds, then $B \\propto \\Sigma_g^{0.85}$. This explanation is more problematic for the B07 relation and $B \\propto \\Sigma_g^{0.7}$. \\footnote{However, we have scaled $B$ to the \\emph{total} Milky Way magnetic field strength of $6~\\muGauss$ near the Solar Circle. If most of the magnetic field strength in starbursts is driven by turbulence, perhaps scaling to the disordered Galactic magnetic field strength \\citep[$\\sim 2 \\muGauss$;][]{Beck01} near the Solar Circle would make more sense, because the ordered magnetic field arises from a different process. The lower normalization for $B$ then would partly compensate for the steeper dependence on $\\Sigma_g$.} However, while we assume that there is a parametrization for $B$ that applied to both compact starbursts and puffy starbursts, the radio excess should arise more generally. The radio excess arises simply because synchrotron cooling time is shorter than the bremsstrahlung and ionization cooling times in puffy starbursts, but longer in compact starbursts at fixed $\\Sigma_g$: at fixed $\\nu^{\\prime}$, $t_{\\rm brems} \\propto n^{-1} \\propto h / \\Sigma_g$ and $t_{\\rm ion} \\propto B^{-1/2} n^{-1} \\propto h / (B^{1/2} \\Sigma_g)$. We therefore expect there to be a radio excess with respect to the FRC in any puffy starburst with a strong enough magnetic field, with the exact enhancement depending on magnetic field strength because of the remaining competition, after ionization and bremsstrahlung are sub-dominant, from the IC losses on starlight.\\footnote{If the magnetic field strength does not go very roughly as $B \\propto \\Sigma_g^{0.7}$, however, there will be more scatter in $L^{\\prime}_{\\rm TIR}/L^{\\prime}_{\\rm radio}$ for puffy starbursts at fixed $h$ and $z$. SMGs would not form their own FRC if $B$ was the same for all SMGs regardless of $\\Sigma_g$, or if $B$ increased \\emph{very} steeply with $\\Sigma_g$, because there still must be a conspiracy with IC losses off starlight. Since SMGs do appear to form their own FRC \\citep[e.g.,][]{Murphy09,Michalowski09}, this could be evidence specifically that their magnetic field strengths increase with $\\Sigma_g$ (or $\\rho$).} Our results imply that a moderate radio excess at the factor of $\\sim 3$ level alone is not a safe indicator of the presence of a radio-loud AGN, especially at high redshifts where SMGs are observed. While radio excess with respect to the local FRC has been suggested as a selection criterion for radio-loud AGNs \\citep[e.g.,][]{Yun01,Yang07,Sajina08}, our models with $\\Sigma_g^{0.7-0.8}$ imply that $q_{\\rm FIR} \\approx 1.7 - 2.0$ for puffy starbursts powered by star-formation alone. A radio excess is inexplicable in our models only when the source is an order of magnitude brighter ($q_{\\rm FIR} \\la 1.5$) in the radio than predicted from the FIR emission. While SMGs are relatively rare and may not be a problem in small samples, we recommend that other means be used to be sure that the radio-excess is caused by an AGN, such as a flat radio spectrum, radio morphology, mid-IR colors, or the presence of strong X-ray emission \\citep[see also][]{Murphy09}. \\subsubsection{$B \\propto \\rho^{0.5 - 0.6}$: Radio-Dim Puffy Starbursts} A magnetic field dependence of $B \\propto \\rho^{0.5}$ appears to hold for Galactic molecular clouds \\citep[e.g.,][]{Crutcher99}, and LTQ found that the FIR-radio correlation was consistent with this magnetic field dependence. The existence of galaxies with different scale heights allows us to distinguish the two possibilities for magnetic field scaling. If $B \\propto \\Sigma_g^a$, then the magnetic field strength will be the same for all galaxies with the same $\\Sigma_g$, regardless of scale height. In models with $B \\propto \\rho^{0.5 - 0.6}$, by contrast, the magnetic field strength is weaker in puffy starbursts than in compact starbursts with the same $\\Sigma_g$. As seen in Figure~\\ref{fig:LFIRRadioRest} (\\emph{right panel}, dotted lines), puffy starbursts again form their own FRC. In models with $B \\propto \\rho^{0.5 - 0.6}$, they are radio dim compared to the $z \\approx 0$ FRC. We show in Table~\\ref{table:Models} that the normalization of the FRC is radio-dim by a factor of $\\sim 1.2 - 2.0$. We can explain this in the LTQ theory of the FRC as well. The magnetic field strength must increase more slowly with $\\rho$ than with $\\Sigma_g$ to reproduce the $z \\approx 0$ FRC, because $h \\propto \\Sigma_g / \\rho$ is 10 times smaller in compact starbursts than normal galaxies and puffy starbursts. Compact starbursts are highly compressed with respect to normal galaxies, so they have strong magnetic fields and synchrotron radio emission is strong enough to compete with the other losses. Puffy starbursts are not compressed, so that their magnetic fields are weak and synchrotron losses cannot keep up with IC losses, nor with bremsstrahlung and ionization as 1.4 GHz emission traces ever lower electron energies at higher magnetic field strengths. Puffy starbursts therefore turn out to be radio dim compared to compact starbursts on the $z \\approx 0$ FRC, if $B \\propto \\rho^{0.5 - 0.6}$. As before, the B07 star-formation relation predicts greater IC losses and therefore weaker radio emission. Because of the claims in the literature that SMGs are radio bright, we do not favor these models. The suggested relative radio brightness of high-$z$ SMGs therefore provides some evidence that, in fact, $B \\propto \\Sigma_g^{0.7 - 0.8}$ rather than $B \\propto \\rho^{0.5 - 0.6}$. However, the matter of whether high-$z$ SMGs are in fact radio bright is not yet settled. Although LTQ concluded that $B$ must increase dramatically from normal galaxies to dense starbursts (\\S~\\ref{sec:Theory}), they were unable to distinguish between these two possibilities with the $z \\approx 0$ FRC alone. For this reason, high-$z$ starbursts and their qualitatively different morphologies compared to those at $z \\approx 0$ can distinguish theories of the FRC. \\subsubsection{Spectral slopes} A prediction of all of our variants is that puffy starbursts like submillimeter galaxies should have steep non-thermal radio spectra, with $\\alpha \\approx 0.8 - 1.0$ (see Table~\\ref{table:Models}). The steep spectra are caused by strong synchrotron cooling in the $B \\propto \\Sigma_g^{0.7-0.8}$ case and the relatively stronger IC cooling off starlight in the $B \\propto \\rho^{0.5 - 0.6}$ case. In general, puffy starbursts should have roughly the same $\\alpha$ as normal galaxies in the local universe, which tends to be somewhat higher ($\\alpha \\approx 0.7 - 1.0$) than in compact starbursts ($\\alpha \\la 0.7$). The slope should hold even out to extremely high $\\Sigma_g$, as long as starbursts are puffy. In contrast, we find that $\\alpha \\approx 0.5$ in compact starbursts, because of efficient ionization and bremsstrahlung losses, which flatten the equilibrium CR spectrum because of their energy dependence. As we note in LTQ, our predicted spectral index for normal galaxies is somewhat too high, and this difference in $\\alpha$ may carry over to the puffy starbursts. However, the significant difference in $\\alpha$ between compact and puffy starbursts should remain as a general prediction of our model: compact starbursts should have flatter spectra than puffy starbursts. The high spectral slopes can be observed either with direct measurements of multifrequency data of individual submillimeter galaxies, or with single frequency observations at a variety of redshifts. There are relatively few measurements of $\\alpha$ for submillimeter galaxies specifically; faint radio sources have $\\alpha \\approx 0.5 - 0.7$ \\citep{Huynh07,Bondi07}, though that sample includes both compact starbursts and AGNs. \\citet{Sajina08} do find that $\\alpha_{610~\\MHz}^{1.4~\\GHz} \\approx 0.8$ for SMGs, comparable to our predictions. They also find that submillimeter galaxies have a radio-excess, in agreement with Figure~\\ref{fig:LFIRRadioRest}. More recently, \\citet{Ibar10} found an average $\\alpha_{610~\\MHz}^{1.4~\\GHz} \\approx 0.75 \\pm 0.06$, which is somewhat flatter than our models. These spectral slopes are not different from normal star-forming galaxies, but are noticeably steeper than local ULIRGs \\citep{Clemens08}. However, we do not account for free-free absorption, which probably flattens the spectra of local ULIRGs like Arp 220 at low frequency \\citep{Condon91}, and is not well understood in SMGs. Since puffy starbursts have steeper spectra than compact starbursts, we expect their inferred $L_{\\rm TIR}^{\\prime}/L_{\\rm radio}^{\\prime}$ will increase with redshift: if the true radio spectral slopes of SMGs are greater than the assumed $\\alpha$ by $\\Delta\\alpha$, they will appear to become radio dimmer by a factor $(1 + z)^{\\Delta \\alpha}$, or up to $\\sim 40\\%$ at $z = 2$. \\begin{figure} \\centerline{\\includegraphics[width=9cm]{f3.eps}} \\figcaption[figure]{The rest-frame synchrotron radio spectra of starbursts with $\\Sigma_g = 1~\\gcm2$, using the B07 star-formation law. Puffy starbursts are dashed, while compact starbursts are solid. These spectra do not include thermal absorption (at low frequencies, $\\la 1\\ \\GHz$) or emission (at high frequencies, $\\ga 30\\ \\GHz$). \\label{fig:StarburstSpectra}} \\end{figure} In Figure~\\ref{fig:StarburstSpectra}, we show the expected radio synchrotron spectra of starburst galaxies, without correcting for thermal absorption or thermal emission. At a rest-frame frequency of 1 GHz, puffy starbursts (dashed) have steeper radio spectra than compact starbursts (solid). Note that at high frequencies ($\\nu^{\\prime} \\ga 10\\ \\GHz$), the ratio of the radio luminosities per unit star formation of the compact and puffy starbursts asymptotes to a value set by the ratio of $U_B$ and $U_{\\rm ph}$ in these starbursts. At these high frequencies, only synchrotron and IC cooling are effective, and IC cooling would be the same for puffy and compact starbursts because of the Schmidt Law (\\S~\\ref{sec:Theory}). For $B \\propto \\Sigma_g^a$, $U_B$ is the same for puffy and compact starbursts, but for $B \\propto \\rho^a$, puffy starbursts have much smaller $U_B$. Thus, measurements of the synchrotron radio emission of SMGs at high $\\nu^{\\prime}$ could determine the magnetic field strength of SMGs and determine which scenario applies. Of course, there are unlikely to be two perfectly distinct populations of compact starbursts and puffy starbursts. Instead, there may be a continuum variation in scale heights from tens to thousands of parsecs. We would then expect to see a larger scatter, both in $L_{\\rm TIR}^{\\prime}/L_{\\rm radio}^{\\prime}$ and $\\alpha$, in a full sample of both the most compact and the most puffy starbursts. \\citet{Murphy09} find that submillimeter galaxies do have a larger scatter in $q_{\\rm TIR}^{\\prime}$ than other galaxies. However, \\citet{Ibar10} find a relatively small scatter of $\\sim 0.3$ in SMG radio spectral index. Importantly, for larger $h$, both $L_{\\rm TIR}^{\\prime}/L_{\\rm radio}^{\\prime}$ and $\\alpha^{\\prime}$ asymptote as CR electron and positron losses are entirely determined by synchrotron and IC; $h \\approx 1~\\kpc$ starbursts are already near this limit. In other words, for arbitrarily large $h$, the radio excess with respect to the $z \\approx 0$ FRC asymptotes to a value of $\\sim 5 - 10$, depending on the assumed Schmidt Law, and the radio spectral slope asymptotes to $\\sim 1.1$ for $p = 2.2$. \\subsection{Synchrotron Radio Emission as a Star-Formation Tracer} \\label{sec:SFTracer} \\begin{figure} \\centerline{\\includegraphics[width=9cm]{f4.eps}} \\figcaption[figure]{The relationship between the inferred rest-frame radio emission and star-formation, using the B07 star formation law, $B \\propto \\Sigma_{\\rm SFR}^{0.7}$, and no winds. Solid lines have $h = 100~\\pc$ for starbursts, while dotted lines have $h = 1~\\kpc$ for starbursts. A flat line would indicate that radio emission is directly proportional to $\\Sigma_{\\rm SFR}$. We see that at low $\\Sigma_g$, the radio flux underestimates the star-formation rate even at $z = 0$, because of CR electron escape (the low-$\\Sigma_g$ conspiracy of \\S~\\ref{sec:Theory}). The radio flux overestimates the star-formation rate for puffy starbursts, because the high-$\\Sigma_g$ conspiracy is unbalanced (\\S~\\ref{sec:Theory}; \\S~\\ref{sec:Puffy}). Finally, the radio emission is suppressed at high redshift, partly because of IC losses off the CMB, and partly because $\\alpha \\ga 0.7$ for puffy starbursts and normal galaxies ($\\alpha = 0.7$ assumed here for the k-correction).\\label{fig:SFRRadio}} \\end{figure} Relying on the FRC, a number of studies have used the GHz radio emission as a tracer of star-formation \\citep[e.g.,][]{Cram98,Mobasher99,Haarsma00,Carilli08,Seymour08,Garn09}. Radio emission has the advantage that it is unaffected by dust obscuration, making it potentially very useful in starbursts, and therefore for most of the star-formation at $z \\ga 1$ \\citep[e.g.,][]{Chary01,LeFloch05,Dole06,Magnelli09,Pascale09}. Our models let us evaluate the theoretical basis for radio as a SFR indicator at all redshifts. We show the predicted radio emissivity as a function of star formation in Figure~\\ref{fig:SFRRadio}. At low surface densities ($\\Sigma_g \\la 0.01~\\gcm2$; $\\Sigma_{\\rm SFR} \\la 0.06~\\Msun~\\kpc^{-2}~\\yr^{-1}$), synchrotron radio has a non-linear dependence on star-formation at all redshifts. The weak radio emission is caused by electrons escaping their host galaxies, as inferred by \\citet{Bell03} and discussed in LTQ. That is, normal galaxies are not perfect electron calorimeters, so the radio emission is not a reliable star-formation tracer at low $\\Sigma_{\\rm SFR}$. At higher redshift, synchrotron emission is diminished by Inverse Compton off the CMB. For a Milky Way-like galaxy, we find that radio is a good star-formation tracer at $z \\la 1$, but underestimates it significantly by $z \\ga 2$ (eq. \\ref{eqn:B07zCrit}). Already galaxies with star formation rates similar to Galactic levels are beginning to be observed in the radio at high redshift \\citep{Garn09}, so that the IC suppression may soon be observed. However, IC losses off the CMB should not be important even in the weakest starbursts until $z \\ga 4$, as is also visible by the redshift evolution of the FRC in Figure~\\ref{fig:FRCEvolution}. Radio emission does grow linearly with star-formation rate between normal galaxies with $\\Sigma_g = 0.01~\\gcm2$ (at $z \\la 2$) and compact starbursts. Therefore, it serves as an acceptable star-formation indicator for these galaxies. However, if $B \\propto \\Sigma_g^{0.7-0.8}$, puffy starbursts like SMGs have about $2 - 4$ times the radio emission at any given star-formation rate than compact starbursts. Therefore, we expect that if $B \\propto \\Sigma_g^{0.7-0.8}$, the usual radio emission estimate based on the $z \\approx 0$ FRC will \\emph{overestimate} their star formation rates by a factor of $\\sim 2 - 4$. The excess is greatest at $\\Sigma_g \\approx 1~\\gcm2$ corresponding to $\\Sigma_{\\rm SFR} \\approx 40 - 200~\\Msun \\kpc^{-2} \\yr^{-1}$, typical of observed SMGs. Among the puffy starbursts themselves, the radio emission grows linearly with star-formation rate. If instead $B \\propto \\rho^{0.5 - 0.6}$, the radio emission will underestimate the star-formation rate. However, because the SMGs lie on their own FRC, there is little real redshift evolution in the radio emissivity of puffy starbursts, because of the buffering provided by IC losses off starlight (see \\S~\\ref{sec:FRCEvolution}, Appendix~\\ref{sec:CMBRadioDim}). Assuming a spectral slope $\\alpha = 0.7$ will also underestimate the radio emissivity, since these galaxies can have steep radio spectra. This explains the apparent evolution with $z$ of puffy starbursts in Figure~\\ref{fig:SFRRadio}: we have applied a k-correction using a typically employed $\\alpha = 0.7$ when, in fact, the true synchrotron spectra are steeper. If the radio star-formation tracer could be calibrated to the special conditions in puffy starbursts like SMGs, taking into account their different scale height, radio emissivity, and spectral slopes, we predict that radio would be a more accurate star-formation tracer for them. \\begin{figure} \\centerline{\\includegraphics[width=9cm]{f5.eps}} \\figcaption[figure]{The \\emph{observed} 1.4 GHz synchrotron radio flux per unit star-formation as a function of redshift, using the B07 star formation law, $B \\propto \\Sigma_{\\rm SFR}^{0.7}$, and no winds. EVLA will have a sensitivity of $\\sim 1\\ \\muJy$ and SKA will have a sensitivity of $\\sim 20\\ \\nJy$.\\label{fig:GHzRadioFluxes}} \\end{figure} We can also directly calculate the \\emph{observed} GHz radio flux density $S_{\\nu}$ from synchrotron emission of star-forming galaxies. We show the predicted flux density per unit star-formation at observer-frame 1.4 GHz in Figure~\\ref{fig:GHzRadioFluxes}. Surveys with the Expanded Very Large Array (EVLA) will have a continuum sensitivity of $\\sim 1\\ \\muJy$ at frequencies of 1 - 50 GHz, and it should be able to directly detect galaxies with $\\Sigma_g \\ga 0.01\\ \\gcm2$ and star-formation rates of $1\\ \\Msun\\ \\yr^{-1}$ out past $z \\approx 0.5$. As stated in \\citet{Murphy09c}, a starburst like M82 with SFR $\\approx 3\\ \\Msun\\ \\yr^{-1}$ will become undetectable past $z \\approx 1$. However, the buffering effect we emphasize in this paper preserves the radio emission of dense starbursts at high $z$, so that bright starbursts will be detectable further: starbursts with SFR $\\ga 100\\ \\Msun\\ \\yr^{-1}$ will be detectable with EVLA at 1.4 GHz out to $z \\approx 4 - 5$ and the most intense starbursts (SFR $\\ga 1000\\ \\Msun\\ \\yr^{-1}$ and $\\Sigma_g \\ga 1~\\gcm2$) will be detectable out past $z \\approx 10$ in synchrotron emission. \\citet{Murphy09c} predicts the Square Kilometer Array (SKA) will be sensitive to star-forming galaxies with flux densities of $\\sim 20\\ \\nJy$. If this sensitivity is attained, then Milky Way-like galaxies ($\\Sigma_g \\approx 0.01\\ \\gcm2$; SFR $\\approx 1\\ \\Msun\\ \\yr^{-1}$) will be directly detectable at 1.4 GHz in synchrotron emission out to $z \\approx 2$, and starbursts with SFR $\\approx 1\\ \\Msun\\ \\yr^{-1}$ will be detectable at 1 GHz beyond $z \\approx 3$. Even at $z \\approx 10$, the synchrotron emission of dense, compact starbursts ($\\Sigma_g \\ga 1\\ \\gcm2$) with SFR greater than $10\\ \\Msun\\ \\yr^{-1}$ should be detectable at 1.4 GHz with SKA (models with $B \\propto \\rho^{0.5-0.6}$ have radio-dim puffy starbursts, and these are detectable at $z = 10$ with the SKA only for SFR greater than $20 - 60\\ \\Msun\\ \\yr^{-1}$). By contrast, \\citet{Murphy09c} found a sensitivity of $25\\ \\Msun\\ \\yr^{-1}$, based on the free-free emission; this limit will apply to normal galaxies and weak starbursts where the synchrotron emission is suppressed. These sensitivities assume that natural confusion, in which radio sources overlap, will not hamper the SKA; estimates for the natural confusion limit vary from nJy to $\\muJy$ levels \\citep[e.g.,][]{Jackson04,Condon09,Murphy09c}. SKA and EVLA will also have good spectral coverage, which may help measurements of the spectral index. If a galaxy is detected at $5 \\sigma$ ($\\sim \\muJy$ for EVLA and $20\\ \\nJy$ for SKA; \\citealt{Murphy09c}) at two different frequencies $\\nu_1$ and $\\nu_2$ with $\\nu_2 / \\nu_1 = 5$, then $\\alpha_1^2$ can be constrained to $\\sim 0.1 - 0.15$ at the $1 \\sigma$ level. EVLA will be better at high observer-frame frequencies (1 - 50 GHz). A problem for the EVLA will be the increasing fraction of thermal emission, which is expected to dominate the emission of starbursts at $\\nu^{\\prime} \\approx 30\\ \\GHz$. EVLA will therefore not easily measure the nonthermal spectral indices of galaxies at high z. On the other hand, the SKA will face free-free absorption when observing low-z starbursts; for example, Arp 220 may be optically thick even at 1 GHz \\citep{Condon91}. At high redshift, however, the rest-frame frequencies SKA will observe will be less affected by free-free absorption. \\subsection{SMGs, The Radio Background, and Radio Source Counts} \\label{sec:RadioBackground} Star-formation in galaxies over cosmic time produces a diffuse radio synchrotron background. Our models indicate that submillimeter galaxies and other puffy starbursts ought to have enhanced radio emission. In principle, this means that the synchrotron radio background could be up to $\\sim 2 - 4$ times higher than usually predicted from the Cosmic Infrared Background (CIB) and a naive application of the $z \\approx 0$ FRC. The magnitude of this implied radio excess is interesting, because ARCADE2 recently reported an excess radio background at 3~\\GHz, about five times higher than expected from star formation \\citep{Fixsen09,Seiffert09}. \\citet{Singal09} found that constraints on Inverse Compton emission require the excess to come from regions with galactic-level ($\\ga \\muGauss$) magnetic fields, and suggest an evolution of the FRC as the source of the reported excess. However, while SMGs are individually very bright and contribute much to the cosmic star-formation rate at $z \\ga 2.5$, they are not typical starbursts \\citep{Bavouzet08}. Instead, they seem to represent a transient phase that can survive about 100~\\Myr~before their gas is depleted \\citep{Tacconi06,Pope08}. We can estimate the total radio background enhancement by scaling to the contribution of SMGs to the CIB, which is largely reprocessed starlight from galaxies at $z \\ga 1$ \\citep[e.g.,][]{Dole06,Devlin09}, and adjusting by the SMG radio excess. Submillimeter galaxies do provide the majority of light at $\\sim 850 \\mum$, but this is only a small fraction of the total IR background. At the peak of the CIB ($\\sim 160 \\mum$), submillimeter galaxies provide $\\la 10\\%$ of the total power, and possibly only $\\sim 2\\%$ \\citep{Chapman05,Dye07}. Optimistically, the radio excess from SMGs would be $10\\% \\times (4 - 1) \\approx 30\\%$. This is significant, but not enough to explain the very large ARCADE2 excess. More conservatively, the excess is more likely $5\\% \\times 2 \\approx 10\\%$, and could be as little as a few percent. Since the number of radio sources down to several $\\muJy$ is well known, we can estimate the fraction of the expected radio background comes from bright SMGs. \\citet{Dole06} find a TIR background of about $24~\\nWBackUnits$; from the normalization of the FIR-radio correlation we expect a 1.4 GHz background of $\\nu I_{\\nu} \\approx 2.6 \\times 10^{-5}~\\nWBackUnits$. \\citet{Chapman05} found an average radio flux density of $S_{1.4} \\approx 75\\ \\muJy$ for bright SMGs ($S_{\\rm 850\\ \\mu m} \\ga 5\\ \\mJy$). The number counts of $75\\ \\muJy$ sources imply that they have a density of $\\sim 1500\\ \\deg^{-2}$, contributing roughly $\\sim 5.1 \\times 10^{-6}~\\nWBackUnits$ to the 1.4 GHz radio background \\citep[e.g.,][]{Gervasi08}. Bright SMGs have an approximate density of $\\sim 600\\ \\deg^{-2}$ \\citep[e.g.,][]{Wang04,Coppin06}, so they constitute about $\\sim 40\\%$ of the background from $75\\ \\muJy$ sources, or $\\sim 8\\%$ of the expected 1.4 GHz background from star-formation. This is roughly in line with our estimate of $\\sim 10\\%$, although the uncertainties are large enough that it could be consistent with SMGs lying on the FRC. In any case, it is fairly clear that bright SMGs are not the source of the ARCADE excess. The total excess could be greater if most starbursts at $z \\ga 1$ are puffy, and not just SMGs. We do not expect this simply because most current studies show that the local FRC \\emph{does} hold out to high redshift for most observed galaxies and starbursts \\citep[e.g.,][]{Murphy09,Younger09,Garn09}. The spectral slope would also be a problem: ARCADE2 inferred $\\alpha = 0.6$, while we predict a spectral slope $\\alpha \\ga 0.7$, steeper than local compact starbursts. Resolved radio sources in the range $50\\ \\muJy \\la S_{1.4} \\la 1\\ \\mJy$, which are expected to be star-forming galaxies \\citep{Danese87,Condon89,Benn93}, contribute about $1.1 \\times 10^{-5}\\ \\nWBackUnits$ to the 1.4 GHz radio background \\citep{Gervasi08}, which is consistent with the total energetics expected from the FRC. \\citet{deZotti10} also find consistency between the FRC and the observed radio source counts at $\\sim 30\\ \\muJy$. Stacking studies of fainter radio sources have given somewhat ambiguous results on whether the FRC applies \\citep{Boyle07,Beswick08}, but \\citet{Garn09b} find no large evolution in the FRC down to $S_{1.4} \\approx 20\\ \\muJy$. Any large radio excess would have to come either from a new population of low luminosity galaxies or very high-$z$ galaxies (see Figure~\\ref{fig:GHzRadioFluxes}); extrapolations of the higher flux source populations do not predict a large radio excess \\citep[e.g.][]{Gervasi08}. Nonetheless, our work indicates that a radio excess from SMGs can be significant. Conversely, if the luminosity function of galaxies was very steep, most galaxies could be intrinsically radio-dim with respect to the FRC, because of IC losses off the CMB (see \\S~\\ref{sec:FRCEvolution} and Figures~\\ref{fig:LFIRRadioRest} and~\\ref{fig:FRCEvolution}). Then the FRC would overestimate the strength of the radio background. However, most star-formation at high $z$ is believed to have occurred in starbursts \\citep{LeFloch05,Dole06}, so this possibility is unlikely. We have applied the theory of LTQ to predict the FRC for redshifts $0 \\le z \\le 10$. We use one-zone models of galaxies and starbursts with CR injection, cooling, and escape to predict the equilibrium, steady-state radio spectra of galaxies and starbursts over the entire range of the FRC. Our goals were to determine how and why the low- and high-$\\Sigma_g$ conspiracies crucial to the $z \\approx 0$ FRC (\\S~\\ref{sec:Theory}) affect the FRC at high redshift, and to provide a quantitative model for predicting the critical redshifts at which galaxies deviate from the $z \\approx 0$ FRC. We find the following: \\begin{enumerate} \\item For compact starbursts ($h \\approx 100~\\pc$), we find relatively little evolution in the FIR-radio correlation out to $z \\approx 5 - 10$ (Figure~\\ref{fig:LFIRRadioRest}). This is partly because the magnetic energy density in galaxies is strong enough to dominate the CMB even at high redshifts. However, the high-$\\Sigma_g$ conspiracy (\\S~\\ref{sec:Theory}) also acts as a buffer against IC losses off the CMB; the increased IC losses must compete with the already present bremsstrahlung, ionization, and IC off starlight in addition to synchrotron losses. The rest-frame radio spectral slope $\\alpha^{\\prime}$ at fixed $\\nu^{\\prime}$ does not change with $z$, but the observed $\\alpha$ at fixed $\\nu$ increases because the non-thermal synchrotron radio spectrum steepens at higher rest-frame frequency. \\item We derive in Appendix~\\ref{sec:CMBRadioDim} the critical redshifts when Inverse Compton losses off the CMB suppress the radio luminosity of galaxies compared to the $z \\approx 0$ FRC. These relations are given for our standard model in equation \\ref{eqn:B07zCrit}. The non-thermal radio luminosity is suppressed severely in Milky Way-like galaxies ($\\Sigma_{\\rm SFR} \\approx 0.06~\\Msun~\\kpc^{-2}~\\yr^{-1}$) at $z \\approx 2$ and the weakest compact starbursts at $z \\ga 5$. The spectrum at GHz steepens to $\\alpha \\approx 1$ because of these enhanced IC losses. Nonetheless, the low-$\\Sigma_g$ conspiracy (\\S~\\ref{sec:Theory}) also acts to prevent the radio emission from steeply falling with redshift, since Inverse Compton losses off the CMB must be more efficient than diffusive escape, not just synchrotron losses (see \\S~\\ref{sec:FRCEvolution}). \\item LTQ found that the $z = 0$ FRC demands that $B$ scales with $\\rho$ or $\\Sigma_g$ in galaxies lying on the Schmidt law. In models with $B \\propto \\Sigma_g^{0.7-0.8}$, we find that puffy starbursts with $h = 1~\\kpc$ such as SMGs are radio bright compared to the $z = 0$ FRC by a factor of $\\sim 2 - 4$. This follows from a breakdown of the high-$\\Sigma_g$ conspiracy (\\S~\\ref{sec:Theory}): bremsstrahlung and ionization cooling are weak in puffy starbursts relative to the compact starbursts that predominate in the $z = 0$ universe. In contrast, in models with $B \\propto \\rho^{0.5 - 0.6}$, we find that puffy starbursts are radio dim compared to the observed FRC, because of weak synchrotron cooling relative to the IC losses. Since several studies have reported radio excesses for SMGs, we favor the $B \\propto \\Sigma_g^{0.7-0.8}$ scaling; however the issue of whether SMGs are radio-bright is still not fully resolved. In either case, puffy starbursts show little true evolution with $z$, though they may appear to have fainter rest-frame radio luminosities at high $z$ because of their steep spectra. Puffy starbursts inevitably have high $\\alpha$ ($\\ga 0.7$), since bremsstrahlung and ionization losses are weak with respect to synchrotron and IC. A key prediction of our scenario with $B \\propto \\Sigma_g^{0.7-0.8}$ is that the variations in $L_{\\rm TIR}^{\\prime}/L_{\\rm radio}^{\\prime}$ will be correlated with scale height at fixed $\\Sigma_{\\rm SFR}$, since the radio-excess in our models is a direct consequence of the large CR scale height and the small bremsstrahlung and ionization losses it causes. Radio-excess (low $q$) puffy starbursts will have steeper radio spectral slopes (bigger $\\alpha$), larger velocity dispersions $\\sigma$ compared to their rotation speeds $v_{\\rm circ}$, and possibly moderately cooler dust temperatures (smaller $T_{\\rm dust}$). \\item As previously expected, radio emission can be a poor tracer of star formation in low surface density galaxies, because of electron escape and IC losses off the CMB. For our preferred $B \\propto \\Sigma_g^{0.7-0.8}$ scaling, radio emission overestimates star-formation rate by a factor of $2 - 4$ in puffy starbursts. Star-formation rate is underestimated by synchrotron radio emission with the $B \\propto \\rho^{0.5 - 0.6}$ scaling. \\item While SMGs may be individually radio bright compared to the local FRC, they contribute a relatively small fraction of the Cosmic Infrared Background and the total star-formation luminosity of the Universe. This means that they enhance the star-formation radio background by $\\la 50\\%$, and possibly around $\\sim 10\\%$, with respect to a naive application of the $z = 0$ FRC. \\end{enumerate} As in LTQ, we did not exactly match the observed radio spectral slopes of galaxies, with $\\alpha \\approx 0.9 - 1.0$ in normal galaxies and $\\alpha \\approx 0.4 - 0.6$ in compact starbursts. This will have a slight effect on the k-correction. An error of $0.25$ in $\\alpha$ should only affect inferred radio luminosity by 30\\% at $z = 2$ and 60\\% at $z = 5$. Nevertheless, the prediction of steeper radio spectra in puffy starbursts with respect to compact starbursts at all relevant $z$ should be robust. Our explanation for the small $L_{\\rm TIR}^{\\prime}/L_{\\rm radio}^{\\prime}$ ratio in submillimeter galaxies as a breakdown of the high-$\\Sigma_g$ conspiracy is based purely on the steady-state spatially-averaged synchrotron emission, but the details of the FIR emission may also matter. Throughout this paper, we have simply assumed that the bolometric FIR luminosity could be correctly inferred from observations, and have assumed the same UV opacity for all galaxies at all redshifts. The total FIR emission is also likely to depend on the metallicity, and may be lower at high $z$ for the lowest surface density galaxies. The exact far-infrared SED is important in determining the FIR emission when observations have only been made at only a few wavelengths. The presence of AGNs, a different IMF at high $z$, and selection biases may also affect the inferred $q$ of SMGs. We did not include the effects of galaxy evolution on the CR spectrum in our models. We argued in LTQ that it should not matter for quiescent spirals or for extreme starbursts, because the CR lifetime is much shorter than the time dependence of stellar populations. However, galaxy evolution may play a role in weaker starbursts \\citep{Lisenfeld96b} and in post-starburst galaxies \\citep{Bressan02}. Studies of merging normal galaxies and galaxies in clusters have indeed found that they are radio bright with respect to the FRC, possibly because of compression of magnetic fields or shock acceleration \\citep{Gavazzi91,Miller01,Murphy09b}. We also assumed that the magnetic field strength at a given density does not depend on redshift. It is not entirely clear how long normal galaxies take to build up their magnetic fields, or even what process is at work \\citep[see the reviews in ][]{Widrow02,Kulsrud08}, though there are theoretical mechanisms that can rapidly generate strong magnetic fields. Studies at $z \\approx 2$ indicate that normal Milky Way-like galaxies had magnetic fields with similar strengths to the present \\citep{Kronberg08,Bernet08,Wolfe08}. At the very highest redshifts, magnetic field strengths might be weaker, because the seed fields were essentially zero compared to the present strengths. Starbursts also may build their magnetic fields up in much shorter times than normal galaxies, and through a different process than normal galaxies \\citep{Thompson09}. Finally, we have used one-zone models, which are appropriate if the CRs sample all of the gas phases in each galaxy's ISM. However, the ISM is known to be clumpy in the Milky Way, in compact starbursts like Arp 220 \\citep[e.g.,][]{Greve09}, and even in the SMGs themselves \\citep{Tacconi06}. A full understanding of the FRC will probably require models that take into account the inhomogeneity of star-forming galaxies." }, "0910/0910.3960_arXiv.txt": { "abstract": "Galactic bars are the most important driver of secular evolution in galaxies. They can efficiently drive gas into the central kiloparsec of galaxies, thus feed circumnuclear starbursts, and possibly help to fuel AGN. The connection between bars and AGN activities has been actively debated in the past two decades. Previous work used fairly small samples and often lacked a proper control sample. They reported conflicting results on the correlation between bars and AGN activity. Here we revisit the bar-AGN and bar-starburst connections using the analysis of bars in a large sample of about 2000 SDSS disk galaxies \\citep*{bar_etal_08}. We find that AGN and star-forming galaxies have similar optical bar fractions, 47\\% and 50\\%, respectively. Both bar fractions are higher than that in inactive galaxies (29\\%). We discuss the implications of the study on the relationship between host galaxies and their central activities. ", "introduction": "Large-scale bars are very common in disk galaxies. At near infrared (NIR) wavelengths, the optical bar fraction averaged across different Hubble types is $\\sim$~60\\% from quantitative bar identification methods, such as the ellipse fitting and structural decomposition \\citep*[e.g.][]{lau_etal_04a, mar_jog_07, men_etal_07, wei_etal_09} , and is $\\sim$~72\\% from visual classification \\citep{esk_etal_00}. At optical wavelengths, quantitative methods yield an average optical bar fraction of 45\\% to 52\\% \\citep*{mar_jog_07, bar_etal_08, agu_etal_09}, while visual classification yields $\\sim$~60\\% \\citep{devauc_63}. The optical bar fraction is somewhat lower than the NIR fraction due to the obscuration of bars by dust lanes and star formation along the bar. Several studies have now moved beyond considering only the bar fraction averaged over many Hubble types. They investigated how the optical bar fraction varies with the Hubble types or host properties. The optical bar fraction rises in galaxies with low Bulge-to-Disk ratio or/and high luminosity \\citep{odewah_96, bar_etal_08, bar_etal_09, agu_etal_09, mar_etal_09}. The non-axisymmetric stellar bar can drive gas from the outer disk to the central kiloparsec, where they trigger star formation. This is supported by several observations, which show that barred galaxies host high gas densities and circumnuclear starbursts \\citep*[e.g.][]{jog_etal_05} and that the central gas concentration is larger in barred than unbarred galaxies \\citep*[e.g.][]{sak_etal_99, she_etal_04}. Due to their efficiency in driving gas inflows in the disks, bars are strong candidates for the triggering of nuclear activities. There is strong evidence for a connection between large-scale bars and circumnuclear starbursts. Barred galaxies show enhanced radio continuum and infrared emissions compared with unbarred ones \\citep[e.g.][]{hummel_81, haw_etal_86}, and starburst galaxies tend to be more barred compared with the non-starburst galaxies \\citep*[e.g.][]{arsena_89, hua_etal_96, ho_etal_97, hun_mal_99}. Bars can also set up resonances, such as the inner Lindblad resonances (ILRs), which can prevent the gas from going further in. Therefore, gas often builds up on a ring (a few hundred parsecs in radius) and the circumnuclear starbursts can occur there \\citep[e.g.][]{jog_etal_05}. The connection between bars and AGNs is less clear (see \\citealt{jogee_06} for a review). Over the last two decades a large number of studies were carried out to identify if such a correlation exists. Most of them compared the bar fraction of the AGN sample with that of a control sample of inactive galaxies. The results are controversial. For example, studies like \\citet{ho_etal_97}, \\citet{hun_mal_99}, \\citet{mul_reg_97}, and \\citet{mar_etal_03} found no excess of bars in Seyfert galaxies, while \\citet*{kna_etal_00}, \\citet{hao_lai_etal_02}, and \\citet*{lau_etal_04b} reported a higher fraction of bars in Seyferts. Previous studies are limited in several aspects. Firstly, the sizes of the samples are often small, including only a few tens of AGNs and control galaxies. In some cases, the control sample was not well matched to the active sample. Secondly, identifications method for bars are not always consistent across samples. Thirdly, the spectral classifications of galaxies as AGNs or inactive galaxies are significantly inconsistent. Most studies adopted the galaxy classifications in NASA/IPAC NED, which could be done by different people using different criteria. Our study tries to overcome some of these limitations by systematically investigating the optical bar fraction of AGNs and non-AGNs in a large number of galaxies from the Sloan Digital Sky Survey (SDSS), using matched active and control samples, and the same consistent quantitative method for identifying bars across samples. ", "conclusions": "With the classification and structural information of $\\sim 2000$ disk galaxies from the SDSS \\citep{bar_etal_08}, we study the optical bar fraction of AGNs, star-forming, and inactive galaxies from the sample. We find that the optical bar fraction of the AGNs is 47\\%, similar to the optical bar fraction of the star-forming galaxies (50\\%). Both are higher than the optical bar fraction of the inactive galaxies (29\\%). This suggests that accurate and consisitent spectral classification is important in evaluating whether there is an excess of bars in AGNs, and could be the reason for controversial results reported in previous studies on the issue. Our study has several imporvements compared to previous ones. The size of our sample is large. We have consistent SDSS spectra for all our galaxies, therefore, we can obtain accurate and consistent spectral classifications for the sample. In addition, the SDSS spectra are taken with the 3\\arcsec aperture, which corresponds to 606 pc to 1.78 kpc in the redshift range of $0.01$30,000 bodies with initial orbits close to those of Jupiter-family comets (JFCs), Halley-type comets, long-period comets, trans-Neptunian objects, and asteroids in the resonances 3/1 and 5/2 with Jupiter, and also of $>$20,000 dust particles produced by these small bodies was integrated during their dynamical lifetimes. We considered the gravitational influence of planets, but omitted the influence of Mercury (exclusive for Comet 2P/Encke). In about a half of calculations of migration of bodies, we used the method by Bulirsh-Stoer (1966) (BULSTO code), and in other runs we used a symplectic method (RMVS3 code). The integration package of Levison \\& Duncan (1994) was used. For dust particles, only the BULSTO code was used, and the gravitational influence of all planets, the Poynting-Robertson drag, radiation pressure, and solar wind drag were taken into account. The ratio $\\beta$ between the radiation pressure force and the gravitational force varied from $\\le$0.0004 to 0.4. For silicates, such values of $\\beta$ correspond to particle diameters $d$ between $\\ge$1000 and 1 microns; $d$ is proportional to 1/$\\beta$. In our calculations, planets were considered as material points, so literal collisions did not occur. However, using the algorithm suggested by Ipatov (1988) with the correction that takes into account a different velocity at different parts of the orbit (Ipatov \\& Mather 2003), and based on all orbital elements sampled with a 10-500 yr step, we calculated the mean probability $P$ of collisions of migrating objects with a planet. The step could be different for different typical dynamical lifetimes of particles. We define $P$ as $P_{\\Sigma}/N$, where $P_{\\Sigma}$ is the probability of a collision of all $N$ objects with a planet. The probabilities of collisions of bodies and particles at different $\\beta$ with planets % are presented in Fig. 1. These probabilities do not take into account the destruction of particles in collisions and sublimation, which can be more important for small particles. Our runs were made mainly for direct modelling of collisions with the Sun, but Ipatov \\& Mather (2007) obtained that the mean probabilities of collisions of considered bodies with planets, lifetimes of the bodies that spent millions of years in Earth-crossing orbits, and other obtained results were practically the same if we consider that bodies disappear when perihelion distance becomes less than the radius of the Sun or even several such radii. ", "conclusions": "" }, "0910/0910.0318.txt": { "abstract": "Collimated supersonic flows in laboratory experiments behave in a similar manner to astrophysical jets provided that radiation, viscosity, and thermal conductivity are unimportant in the laboratory jets, and that the experimental and astrophysical jets share similar dimensionless parameters such as the Mach number and the ratio of the density between the jet and the ambient medium. When these conditions apply, laboratory jets provide a means to study their astrophysical counterparts for a variety of initial conditions, arbitrary viewing angles, and different times, attributes especially helpful for interpreting astronomical images where the viewing angle and initial conditions are fixed and the time domain is limited. Experiments are also a powerful way to test numerical fluid codes in a parameter range where the codes must perform well. In this paper we combine images from a series of laboratory experiments of deflected supersonic jets with numerical simulations and new spectral observations of an astrophysical example, the young stellar jet HH~110. The experiments provide key insights into how deflected jets evolve in 3-D, particularly within working surfaces where multiple subsonic shells and filaments form, and along the interface where shocked jet material penetrates into and destroys the obstacle along its path. The experiments also underscore the importance of the viewing angle in determining what an observer will see. The simulations match the experiments so well that we can use the simulated velocity maps to compare the dynamics in the experiment with those implied by the astronomical spectra. The experiments support a model where the observed shock structures in HH 110 form as a result of a pulsed driving source rather than from weak shocks that may arise in the supersonic shear layer between the Mach disk and bow shock of the jet's working surface. ", "introduction": "Collimated supersonic jets originate from a variety of astronomical sources, including active galactic nuclei \\citep{agnref}, several kinds of interacting binaries \\citep{binaryjetref}, young stars \\citep{ysojetref}, and even planetary nebulae \\citep{pnjetref}. Most current jet research focuses on how accretion disks accelerate and collimate jets, or on understanding the dynamics of the jet as it generates shocks along its beam and in the surrounding medium. Both areas of research have broad implications for astrophysics. Models of accretion disks typically employ magnetized jets to remove the angular momentum of the accreting material. The distribution and transport of the angular momentum in an accretion disk affects its mass accretion rate, mixing, temperature profile, and density structure, and in the case of young stars, also helps to define the characteristics of the protoplanetary disk that remains after accretion ceases. At larger distances from the source, shock waves in jets clear material from the surrounding medium, provide insights into the nature of density and velocity perturbations in the flow, and enable dynamical studies of mixing, turbulence and shear. Jets from young stars are particularly good testbeds for investigating all aspects of the physics within collimated supersonic flows \\citep[see][for a review]{ray07}. Shock velocities within stellar jets are low enough that the gas cools by radiating emission lines rather than by expanding. Relative fluxes of the emission lines determine the density, temperature, and ionization of the postshock gas, while the observed Doppler shifts and emission line profiles define the radial velocities and nonthermal motions within the jet \\citep[see][for a review]{hartigan08}. Moreover, many stellar jets are located relatively close to the Earth, so that one can observe proper motions in the plane of the sky from observations separated by several years \\citep{heathcote92}. Combining this information with radial velocity measurements gives the orientation of the flow to the line of sight. Using the Hubble Space Telescope, one can observe morphological changes of knots within jets, and follow how these changes evolve in real time \\citep{hartigan01}. Results from these studies show that internal shock waves, driven by velocity variations in the flow, sweep material in jets into a series of dense knots. Typical internal shock velocities are $\\sim$ 40 km$\\,$s$^{-1}$, or $\\sim$ 20\\%\\ of the flow speed. In several cases it is easy to identify both the bow shock and the Mach disk from emission line images. Because stellar jets are mostly neutral, strong H$\\alpha$ emission occurs at the shock front where neutral H is collisionally excited \\citep{heathcote96}. Forbidden line radiation occurs in a spatially-extended cooling zone in the postshock material. Temperatures immediately behind the internal shocks can exceed $10^5$K, but the gas in the forbidden-line-emitting cooling zone is typically 8000~K. Young stars show a strong correlation between accretion and outflow, leading to the idea that accretion disks power the outflows \\citep{heg95,cabrit07}. Most current models use magnetic fields in the disk to launch a fraction of the accreting material from the disk into a collimated magnetized jet \\citep{ferreira06}. Stellar jets often precess, and there is some evidence that they rotate \\citep{ray07}, although rotation signatures are difficult to measure because the rotational velocities are typically only a few percent of the flow speeds and precession can mimic rotational signatures \\citep{cerquieriaref}. In all cases proper motion measurements show that jets move radially away from the source. Jets show no dynamical evidence for kink instabilities, and in fact while magnetic fields may dominate in the acceleration regions of jets, they appear to play a minor role in the dynamics at the distances where most jet knots are observed \\citep{hartigan07}. In the cooling zones behind the shock waves the plasma $\\beta$ can drop below unity, so that magnetic fields dominate thermal pressure in those areas. However, the magnetic pressure is small compared with the ram pressure of the jet. While more unusual than internal working surfaces produced by velocity perturbations, shocks also occur when jets collide with dense obstacles such as a molecular cloud. When the obstacle is smaller than the jet radius, it becomes entrained by the jet, and a reverse bow shock or `cloudlet shock' forms around the obstacle \\citep{schwartz78,l1551}. Alternatively, when a large obstacle like a molecular cloud deflects the jet, a quasi-stationary deflection shock at the impact point forms, followed by a spray of shocked jet material downstream. The classic example of such a jet is HH 110 \\citep{riera03,lopez05}. Though the observations summarized above provide a great deal of information about stellar jets, several important questions remain unanswered. The basic mechanism by which disks load material onto field lines (assuming the MHD disk scenario is correct), and the overall geometry of this wind is unknown, and the roles of reconnection and ambipolar diffusion in heating the jet close to the star are unclear. At larger distances, the magnetic geometry and its importance in shaping the internal working surfaces is poorly-constrained, as are the time scales and spatial scales associated with mixing in supersonic shear layers and working surfaces. The inherently clumpy nature of jets also affects the flow dynamics and observed properties of jets in uncertain ways, and the degree to which fragmentation and turbulence influence the morphologies of jets is unknown. Developing laboratory analogs of stellar jets could help significantly in addressing the questions above. Observations of a specific astronomical jet are restricted to a small range of times and to a particular observing angle, while laboratory experiments have no such restrictions. In principle, one could explore a wide range of initial collimations, velocity and density structures within the jet, as well as densities, geometries, and magnetic field configurations with laboratory experiments. The experiments also provide a powerful and flexible way to test 3-D numerical fluid codes, and to investigate how real flows develop complex morphologies in 3-D. The challenge is to design an experiment that is relevant to the astrophysical case of interest. Laboratory experiments differ by 15 $-$ 20 orders of magnitude in size, density and timescale from stellar jets, but because the Euler equations that govern fluid dynamics involve only three variables, time, density, and velocity, the behavior of the fluid is determined primarily by dimensionless numbers such as the Mach number (supersonic or subsonic), Reynolds number (viscous or inertial), and Peclet number (importance of thermal conduction). If the experiments behave as a fluid and have similar dimensionless fluid numbers as those of stellar jets, then the experiments should scale well to the astrophysical case \\citep{ryutov99}. Other parameters, such as magnetic fields and radiational cooling are more difficult to match, and it is impossible to study the non-LTE excitation physics in the lab because the critical density for collisional deexcitation is not scalable. In any case, the materials are markedly different between the experiments and stellar jets, so it is not possible to study emission line ratios in any meaningful way with current laboratory capabilities. Hence, at present the main utility of laboratory experiments of jets is to clarify how complex supersonic flows evolve with time. Laboratory work relevant to stellar jets is an emerging area of research, and several papers have appeared recently which address various aspects of supersonic flows in the appropriate regime \\cite[see][for a review]{remington06}. \\citet{hansen07} observed how a strong planar shock wave disrupts a spherical obstacle and tested numerical models of the process, and \\citet{loupias} generated a laboratory jet with a Mach number similar to that of a stellar jet. In a different approach, \\citet{lebedev04} and \\citet{ampleford07} used a conical array of wires at the Magpie facility to drive a magnetized jet, and explored the geometry of the deflection shock formed as a jet impacts a crosswind. Laboratory experiments have also recently studied the physics associated with instabilities along supernova blast waves \\citep{snrref,drake09}, and the dynamics within supernovae explosions \\citep{snrcoreref}. In this paper we present the results of a suite of experiments which deflect a supersonic jet from a spherical obstacle, where the dimensionless fluid parameters are similar to those present in stellar jets. In section 2, we describe our experimental design, consider how the experiments scale to astrophysical jets, demonstrate that the experiments are reproducible, and report how the observed flows change as one varies several parameters, including the distance between the axis of the jet and the obstacle, the time delay, the density probed with different backlighters, and the viewing angle. The numerical work is summarized in section 3. Detailed calculations with the 3-D RAGE code reproduce all of the major observed morphologies well. In section 4, we present new high-resolution optical spectra of the shocked wake of the HH 110 protostellar jet, and a new wide-field H$_2$ image of the region. These observations quantify how the internal dynamics of the gas behave as material flows away from the deflection shock and show how the jet entrains material from the molecular cloud core. Finally, in section 5 we consider how the experiments and the simulations from RAGE provide new insights into the internal dynamics of deflected supersonic jets, especially in the regions of the working surface and at the interface where jet material entrains and accelerates the obstacle. ", "conclusions": "%6 The combination of experimental, numerical and astronomical observational data from this study demonstrates the potential of the emerging field of laboratory astrophysics. In this paper we have studied how a supersonic jet behaves with time as it deflects from an obstacle situated at various distances from the axis of the jet. The laboratory analog of this phenomenon scales very well to the astrophysical case of a stellar jet which deflects from a molecular cloud core. An important component to our study was to expand the observational database of best astrophysical example, HH~110, by obtaining new spatially-resolved high-spectral resolution observations capable of distinguishing thermal motions from turbulent motions, and by acquiring a new deep infrared H$_2$ image that can be compared with existing optical emission line images from the Hubble Space Telescope. The laboratory experiments span a range of times, spatial offsets between the axis of the jet and the center of the ball (impact parameters), viewing angles, opacities (backlighters), and materials. The experiments are reproducible and do not depend on composition or structure of foam, or on the pinhole diameter of the backlighter (spatial resolution of the experiment). Synthetic radiographs of the experiment from RAGE match the experimental data extremely well, both qualitatively as images and quantiatively with Fourier analysis. In fact, the agreement is so good that we have used the synthetic velocity maps from RAGE to compare the internal dynamics of the experiment with those that we measure from the new spectral maps of HH 110. A new wide-field H$_2$ image supports a scenario where HH~110 represents the shocked `spray' that results from a glancing collision of the HH~270 jet with a molecular cloud core. The H$_2$ in HH~110 is offset from the H$\\alpha$ toward the side closest to the molecular core, consistent with the deflected jet model. The H$_2$ images also uncovered two sources within the core that drive collimated jets, one a bright near-infrared source, and the other a highly-obscured source that appears to be a dense protostellar disk observed nearly edge-on, as in HH~30. The experiments provide several important insights into how deflected supersonic jets like HH~110 behave. In the experiment, entrainment of material in the obstacle occurs in part because the morphology of the contact discontinuity between the shocked jet and shocked obstacle easily develops a complex 3-D structure of cavities that enables the jet to isolate clumps of obstacle material and entrain them into the flow. A similar process likely operates in HH~110. The experiments also reveal filamentary structure in the working surface area of the deflected bow shock, but the relative motion between these filaments is subsonic. Hence, while this dynamical process will generate density fluctuations in the outflowing gas, it cannot produce the filamentary structure and $\\sim$ Mach 5 shocks shown by the new velocity maps of HH~110. For this reason the best model for HH~110 remains that of a pulsed jet which interacts with a molecular cloud core. Synthetic position-velocity maps along the deflected jet from the RAGE simulations of the experiments appear as a series of arcs, similar to those observed in astronomical observations of resolved bow shocks. A close examination of the experimental data shows that these arcs correspond to regions where the slit crosses different regions of the flow, such as cavities evacuated by the jet, the jet itself, or entrained material from the ball. This correspondance between the appearance of a p-v diagram and the actual morphology of a complex flow is an intuitive, although perhaps unexpected result of studying the dynamics within the experimental flow. Finally, observations of the deflected bow shock from different viewing angles emphasize that the observed morphology and collimation properties of bow shocks depend strongly upon the orientation of the flow with respect to the observer. As one would expect, a bow shock deflected toward the observer appears less collimated than one that is redirected into the plane of the sky. The impact parameter of the jet and obstacle determines how much the jet deflects from the obstacle and how rapidly the obstacle becomes disrupted by the jet. While experimental analogs of astrophysical jets are highly unlikely to ever reproduce accurate emission line maps, laser experiments can provide valuable insights into how the dynamics of complex flows behave. Our study of a deflected supersonic jet is only one example of how the fields of astrophysics, numerical computation, and laboratory laser experiments can compliment one another. We are currently embarking on a similar program to study the dynamics within supersonic flows that are highly clumpy, and other investigations are underway related to the launching and collimation of jets \\citep{bellan09}." }, "0910/0910.5227_arXiv.txt": { "abstract": "We have studied the effects of electron--ion non-equipartition in the outer regions of relaxed clusters for a wide range of masses in the $\\Lambda$CDM cosmology using one-dimensional hydrodynamic simulations. The effects of the non-adiabatic electron heating efficiency, $\\beta$, on the degree of non-equipartition are also studied. Using the gas fraction $f_{\\rm gas} = 0.17$ (which is the upper limit for a cluster), we give a conservative lower limit of the non-equipartition effect on clusters. We have shown that for a cluster with a mass of $M_{\\rm vir}\\sim 1.2 \\times 10^{15} M_{\\odot}$, electron and ion temperatures differ by less than a percent within the virial radius $R_{\\rm vir}$. The difference is $\\approx 20\\%$ for a non-adiabatic electron heating efficiency of $\\beta \\sim 1/1800$ to $0.5$ at $\\sim 1.4 R_{\\rm vir}$. Beyond that radius, the non-equipartition effect depends rather strongly on $\\beta$, and such a strong dependence at the shock radius can be used to distinguish shock heating models or constrain the shock heating efficiency of electrons. With our simulations, we have also studied systematically the signatures of non-equipartition on X-ray and Sunyaev--Zel'dovich (SZ) observables. We have calculated the effect of non-equipartition on the projected temperature and X-ray surface brightness profiles using the MEKAL emission model. We found that the effect on the projected temperature profiles is larger than that on the deprojected (or physical) temperature profiles. The non-equipartition effect can introduce a $\\sim 10\\%$ bias in the projected temperature at $R_{\\rm vir}$ for a wide range of $\\beta$. We also found that the effect of non-equipartition on the projected temperature profiles can be enhanced by increasing metallicity. In the low-energy band $\\lesssim 1$~keV, the non-equipartition model surface brightness can be higher than that of the equipartition model in the cluster outer regions. Future X-ray observations extending to $\\sim R_{\\rm vir}$ or even close to the shock radius should be able to detect these non-equipartition signatures. For a given cluster, the difference between the SZ temperature decrements for the equipartition and the non-equipartition models, $\\delta\\Delta T_{\\rm SZE}$, is larger at a higher redshift. For the most massive clusters at $z \\approx 2$, the differences can be $\\delta\\Delta T_{\\rm SZE} \\approx $ 4--5$~\\mu$K near the shock radius. We also found that for our model in the $\\Lambda$CDM universe, the integrated SZ bias, $ Y_{\\rm non{\\text -}eq} /Y_{\\rm eq}$, evolves slightly (at a percentage level) with redshift, which is in contrast to the self-similar model in the Einstein--de Sitter universe. This may introduce biases in cosmological studies using the $f_{\\rm gas}$ technique. We discussed briefly whether the equipartition and non-equipartition models near the shock region can be distinguished by future radio observations with, for example, the Atacama Large Millimeter Array. ", "introduction": "\\label{sec:intro} Observational and theoretical studies have shown that the study of the intracluster medium (ICM) can be used as a test of plasma physics under extreme environments that cannot be achieved in terrestrial laboratories, as well as an important cosmological probe. If we assume the matter content of clusters is a fair sample of the universe, the baryon fraction of clusters can be used as an estimator of the average value for the universe, with proper correction for the baryons contained in the stellar component and for a small amount of baryonic matter expelled from clusters during the formation process. Recent work from \\citep{ARS08} has shown that the combined results of the baryon fraction from X-ray observation with the cosmic microwave background (CMB) data can give powerful constraints on cosmological parameters, such as the equation of state of the dark energy. However, the study of cosmology using clusters of galaxies relies heavily on the understanding of cluster physics. For precision cosmology, systematic uncertainties at even the percent level are significant. For example, in current studies, the baryon fraction within clusters is assumed to be independent of redshift and the mass of clusters. It would be important to see if these assumptions are justified; if not, it is important to study the dependence of the baryon fraction on cluster properties and redshift. Even if the dependence on redshift is weak, the correction factor for the baryon content within clusters compared to the average value in the universe could affect the constraints on cosmological parameters. Studying cluster outskirts ($\\gtrsim R_{200}$\\footnote{ $R_{\\Delta}$ is the radius within which the mean total mass density of the cluster is $\\Delta$ times the critical density. The virial radius $R_{\\rm vir}$ is defined as a radius within which the cluster is virialized. For the Einstein--de Sitter universe, $ R_{\\rm vir} \\approx R_{178}$, while for the standard $\\Lambda$CDM universe, $ R_{\\rm vir} \\approx R_{95}$. }) is very important because the boundary conditions of the cluster outskirts constrain the global properties of a cluster \\citep[e.g.,][]{TSN00}. Also, the outer envelopes of clusters have been thought to be less subject to some additional physics including active galactic nucleus (AGN) feedback, and that the outer regions of clusters may provide better cosmological probes. Currently, there are very few observations of the properties of the ICM in the outer parts of clusters. Thus, most of our understanding of these regions is still based on numerical hydrodynamic simulations which assume the hot plasma is a fluid. In these simulations, the clusters are formed from mergers and accretion of dark matter and baryonic gas in overdense regions. A variety of shocks with different geometries along the large-scale structure (LSS) filaments and transverse to them near and beyond the virial radius are unambiguous predictions of the cosmological hydrodynamic simulations. Unfortunately, the lack of observational information on the clusters outskirts prevents us from understanding the accretion shock region, and hence the input physics for the numerical simulations is called into question. For example, the thermodynamic state of the shocked gas, as well as the shock position, depends on the pre-shock gas temperature; the shock will be weaker if the infalling gas is pre-heated \\citep{TSN00}. Even worse, recently it was noted that the non-fluid properties may be important in regions near the virial radius, where the Coulomb collisional mean free path is comparable to the cluster size of a few Mpc \\citep{Loe07}. The Coulomb collisional timescale can also be of the order of the age of the cluster. This suggests that a full kinetic gas theory is needed instead of the fluid approximation when studying the gas properties near the edge of the cluster. Direct consequences include non-equipartition between electrons and ions \\citep{FL97, EF98}, element sedimentation \\citep{CL04}, and suprathermal evaporation of hot gas from the clusters (\\citealt{Loe07}; see also \\citealt{Med07}). Some of these effects can lead to a bias in baryon fraction measurements, and hence cosmological studies. Recently, progress has been made in the study the baryon content of the outer regions of clusters through the X-ray observations together with the Sunyaev--Zel'dovich (SZ) effect on the CMB by the hot electrons in the ICM out to $\\sim R_{200}$ \\citep{ALN+07}. A 3$\\sigma$ result from the {\\it WMAP} three-year data suggests that $35\\%\\pm8\\%$ of the thermal energy in ICM are missing, indicating that the baryons in clusters may be missing even accounting for those locked in stars. The result is also supported by independent measurements from other X-ray and SZ observations \\citep{Ett03, LBC+06, VKF+06, Evr+08}. Using the X-ray observations together with numerical simulations, \\citet{Evr+08} reported that as much as $50\\%$ of the thermal energy can be missing in the ICM. Although \\citet{Gio+09} reported that the total baryon fraction within a smaller radius of $R_{500}$ of massive clusters are consistent to the cosmic value within $1 \\sigma$ when all the X-ray hot gas, stellar mass in galaxies, gas depletion during cluster formation, and intracluster light from stars are taken into account, if the missing baryons measured in the outer region ($\\lesssim R_{200}$) is really significant, this may indicate either a yet-unknown baryonic component, or some new astrophysical processes in the ICM which is driving out the gas from the clusters. While there is no evidence for any undetected baryonic component, \\citet{ALN+07} pointed out that the missing of hot baryons can either be explained by the thermal diffusion or the evaporation of baryons out of the virial radius of clusters. Another possibility is that electron temperature is lower than that of the equipartition value \\citep{WSL+08}. Given the advancements in the X-ray and SZ observations of the cluster outer regions \\citep{ALN+07, Bau+09, GFS+09, Rei+09}, as well as the growing evidence of missing thermal energy in the ICM and the possible negative implications for cosmological tests, a more detailed study of the kinetic processes in cluster envelopes is necessary. While magnetic fields may affect some of the kinetic effects of transport processes such as thermal conduction, the magnetic effects on non-equipartition should not be important since the physics is local. Moreover, it is known that various astrophysical shocks in magnetized environment lead to non-equipartition \\citep{GLR07, HSL+01}. The collisionless accretion shock at the outer boundary of a cluster should primarily heat the ions since they carry most of the kinetic energy of the infalling gas. Assuming that cluster accretion shocks are similar to those in supernova remnants, the electron temperature, $T_e$, immediately behind the shock would be lower than the ion temperature, $T_i$. The equilibration between electrons and ions would then proceed by Coulomb collisions. Near the virial radius, due to the low density, the Coulomb collisional timescale can be comparable to the age of the cluster, and the electrons and ions may not achieve equipartition in these regions \\citep{FL97}. Since X-ray and SZ observations measure the properties of the electrons in the ICM, the net effect is to underestimate the total thermal energy content within clusters. This might account for some or all of the missing thermal energy in the ICM derived by the X-ray and SZ observations. As mentioned above, non-equipartition of ions and electrons is observed in various astrophysical shocks. Most supernova remnants with high Mach numbers comparable to cluster accretion shocks have electron temperatures which are lower than the ion temperatures \\citep{GLR07}; in situ measurements from satellites show the same feature in the Earth's bow shock \\citep{HSL+01}. On the other hand, X-ray observations of the merger shock in the Bullet Cluster indicate that the equilibration time may be shorter than that expected from Coulomb collisions alone \\citep{MV07}. However, this merger shock has a Mach number of a few. From both supernova remnant measurements and a physical model, \\citet{GLR07} have shown that electron heating efficiency within a shock front (usually tens of the gyroradius) is inversely proportional to the Mach number squared. If the results can also be applied to the ICM, the low Mach number merger shocks would be immediately heated to $T_e/T_i \\sim 1$, while cosmological accretion shocks with much higher Mach numbers would only be heated to $T_e/T_i \\ll 1$ by collisionless processes. After the electrons and ions pass through the thin shock front, they will likely be equilibrated by Coulomb collisions alone \\citep{BDD08, BPP08}. The non-equipartition in cluster of galaxies has been previously studied by \\citet{FL97} and \\citet{EF98} in semianalytic models. They have shown that the temperature difference can be significant in the outer one-third of the shock radius of a cluster. One- or three-dimensional simulations for some individual clusters have also been studied \\citep{CAT98, Tak99, RN09}. While these simulations use different cluster or cosmological models, a general agreement is that the effect of non-equipartition is important if shock heating efficient of electrons is low ($\\ll 1$) and the equilibration afterward is due to Coulomb collisions alone. In this paper, we study systematically the effects of non-equipartition on X-ray and SZ observables in outer regions of relaxed galaxy clusters, which is particularly important for cosmological studies. We carry out one-dimensional hydrodynamic simulations with realistic Navarro-Frenk-White (NFW) model under the concordance $\\Lambda$CDM cosmological background to provide a sample of clusters (groups) with different masses ($10^{13}$--$10^{16} M_{\\odot}$) at different redshifts ($z=0$--$2$). Even though we are studying the kinetic non-fluid properties in the cluster outer regions, the hydrodynamic treatment in modeling the cluster dynamical properties is reasonable and is justified as follows. Even dynamically unimportant magnetic fields should be able to reduce significantly the diffusion mean free path perpendicular to the magnetic field \\citep{Sar86,BK09}. The suppression of diffusion in a plasma depends on the topology of magnetic fields. For uniform magnetic fields, only diffusion perpendicular to the local magnetic field is suppressed, and along the field, particles move freely; their mean free path along a field line is still determined by Coulomb collisions. On large scales, diffusion is suppressed in bulk only if the magnetic fields are random and highly tangled on small scales. To include anisotropic diffusion in the calculation would be difficult since the magnetic field structure is not known well enough. However, there is some evidence from large-scale magnetohydrodynamic simulations that magnetic fields in galaxy clusters are chaotic with correlation and reversal length scales of $\\sim 50$ and $\\sim 100$~kpc, respectively \\citep{DBL02}. Hence, we simply assume that diffusion is suppressed. We also assume that electrons and ions are equilibrated locally on a long Coulomb collisional timescale, and assume that equilibration via plasma instabilities \\citep{SCK+05, SCK+08} does not occur except at the shocks (see Section~\\ref{sec:e-heating}). Previous studies show that the dynamical properties of cluster outer regions in one-dimensional simulations successfully reproduce those simulated in three-dimensional calculations \\citep{NFW95, RK97}. The advantages of the one-dimensional simulations for our problem are presented in Section~\\ref{sec:hydro}. We emphasize the signatures of non-equipartition on X-ray and SZ observations in our studies. We also study the effect of electron shock heating efficiency on the degree of non-equipartition. Thus, observations of electron--ion equilibration may give constraints to the electron heating efficiency, and hence the electron heating mechanism. The wider parameter space compared to previous work explored in this paper allows us to study the impact of non-equipartition effects on cosmological studies in a future paper. The paper is organized as follows. In Section~\\ref{sec:hydro}, we describe the set up of our hydrodynamic models. The detailed implementations of the shock heating and the Coulomb equilibration process for our simulations are presented in Section~\\ref{sec:e-heating}. The ability of our simulations to reproduce analytic test models relevant to our studies is discussed in Section~\\ref{sec:test}. We present the simulated dynamics of our realistic NFW cluster models in the standard $\\Lambda$CDM cosmology in Section~\\ref{sec:NFW-DE_dynamics}. These cluster models are used to study the non-equipartition effects presented in the paper. We define the X-ray and SZ observables for our models to be studied, and also present the results for these observables in Section~\\ref{sec:obs}. We discuss and conclude our work in Section~\\ref{sec:conclusion}. Unless otherwise specified, we assume the Hubble constant $H_0 = 71.9~h_{71.9}$~km~s$^{-1}$~Mpc$^{-1}$ with $h_{71.9}=1$, the total matter density parameter $\\Omega_{M,0} = 0.258$, the dark energy density parameter $\\Omega_{\\Lambda} = 0.742$, and the gas fraction $f_{\\rm gas}=\\Omega_b/\\Omega_M=0.17$, where $\\Omega_b$ is the baryon density parameter, for the realistic NFW model in the standard $\\Lambda$CDM cosmology\\footnote{\\tiny http://lambda.gsfc.nasa.gov/product/map/dr3/parameters\\_summary.cfm}, and a hydrogen mass fraction $X=76\\%$ for the ICM throughout the paper. ", "conclusions": "\\label{sec:conclusion} Using one-dimensional hydrodynamic simulations, we have calculated a sample of realistic NFW clusters in a range of masses in the $\\Lambda$CDM cosmology. The cluster properties we simulated are consistent with the one-dimensional $N$-body simulations by \\citet{RK97}, and they have shown that their calculations reproduce the density and temperature profiles of the three-dimensional simulated relaxed clusters in the outer regions. Our one-dimensional hydrodynamic simulations help us to isolate the important physical processes under controlled conditions. We have studied in detail the effect of non-equipartition in the outer regions of relaxed clusters in the $\\Lambda$CDM cosmology. Using $f_{\\rm gas} = 0.17$ (which is the upper limit for a cluster), we give a conservative lower limit of the non-equipartition effect on clusters. We have shown that for a cluster with a mass of $M_{\\rm sh}\\sim 1.5 \\times 10^{15} M_{\\odot}$, within $R_{\\rm vir}$, electron and ion temperatures only differ by less than a percent. Our results show that the effect is smaller than those calculated from recent three-dimensional simulations, which shows that $T_e$ can be biased low by 5\\% at $R_{200} \\sim 0.7 R_{\\rm vir}$ \\citep[model CL104 in][]{RN09}. A detailed analysis is needed to address the difference, but a possible explanation may be that in a three-dimensional cluster the accretion shock can be formed further in. Our results show that $T_e/{\\bar T}$ can reach $\\approx 0.8$ for a range of non-adiabatic electron heating efficiency $\\beta \\sim 1/1800$ to $0.5$ at $\\sim 0.9 R_{\\rm sh}$ (or $\\sim 1.4 R_{\\rm vir}$). Beyond that radius, $T_e/{\\bar T}$ depends rather strongly on $\\beta$, and such a strong dependence at the shock radius can be used to distinguish shock heating models or constraint the shock heating efficiency of electron. We also show that the effect of non-equipartition is larger for more massive clusters, which is consistent to analytic self-similar models in the Einstein--de Sitter universe \\citep{FL97}. Using the algorithm developed by \\citet{Vik06} which takes into account the soft emission at low temperature down to $\\sim 0.5$~keV, arbitrary metallicity, and instrumentation response, we calculated the X-ray spectroscopic-like temperature profiles which are the one to be directly determined observationally. The effect of non-equipartition on the projected temperature profiles is larger than that on the deprojected (or physical) temperature profiles. Non-equipartition effects can introduce a $\\sim 10\\%$ bias in the projected temperature at $R_{\\rm vir}$ for a wide range of $\\beta$. This is because for the projected temperature profile, electrons in the outer region also contribute to the inner region. We also found that the effect of non-equipartition on the projected temperature profiles can be enhanced by increasing metallicity. This is because the domination of the line emissions in the soft band spectra is enhanced by increased metallicity, which is more important for the non-equipartition model where the electron temperature is lower. The effect of non-equipartition on X-ray surface brightness profile in the 0.3--2~keV band is smaller than that on the projected temperature profile. This is because the surface brightness depends on density squared but with a weaker dependence on temperature. This means that in the outer regions, clusters in non-equipartition have similar X-ray surface brightness profile for $E \\lesssim 2$~keV, but with bigger difference in temperature compared to those equipartition counterparts. For $E \\gtrsim 2$~keV, non-equipartition effects on X-ray surface brightness profiles are similar to those on the projected temperature profiles. For a cluster with $M_{\\rm sh}\\sim 1.5 \\times 10^{15} M_{\\odot}$, the effect of non-equipartition on surface brightness profiles in all energy bands is $\\lesssim 10\\%$ for radii $\\lesssim 3$~Mpc; beyond that, the effect can be important. We found that for the non-equipartition model, the surface brightness profile in the low-energy band $\\lesssim 1$~keV can be higher than that of the equipartition model in the cluster outer regions. Non-equilibrium ionization, which was not considered in our calculations of the emissivities, may further enhance the line emissions in the soft bands ($E \\lesssim 1$~keV). Current X-ray observations extend to only $\\sim R_{200} \\sim 2$~Mpc, although some results from recent $Suzaku$ observations begin to go a bit beyond that \\citep{GFS+09}. Within those regions with $\\lesssim R_{200}$, electrons and ions should be almost in equipartition and the signatures in the X-ray temperature and surface brightness should be rather weak. But future X-ray observations may extend to $\\sim R_{\\rm vir} \\approx 1.4 R_{200}$ or even close to the shock radius. We have shown that non-equipartition of electrons and ions should be detectable in those studies. The results by \\citet{GFS+09} support our conclusion. The effects of non-equipartition on the SZ effect were studied. At $z=0$, the effect on the Comptonization parameters is similar to that of the projected temperature profiles. For a cluster with $M_{\\rm sh}\\sim 1.5 \\times 10^{15} M_{\\odot}$, $y_{\\rm non{\\text -}eq}/y_{\\rm eq} \\approx 0.93 (0.8)$ at $1 (1.3) R_{\\rm vir}$. For a given cluster, the difference between the SZ temperature decrements for the equipartition and the non-equipartition models is larger at a higher redshift. For the most massive clusters at $z \\approx 2$, the differences can be $\\delta\\Delta T_{\\rm SZE} \\approx$~4--5$~\\mu$K near the shock radius. A detailed analysis of whether the equipartition and non-equipartition models near the shock region can be distinguished by, for example, ALMA, will be presented in a future paper. The effects on the integrated SZ Comptonization parameter, which measures the thermal energy content of the electrons, were studied. We have shown that the integrated SZ bias, $Y_{\\rm non{\\text -}eq}/Y_{\\rm eq}$, increases as the cluster mass increases, which is expected as the effect of non-equipartition increases with mass. In general, the non-equipartition effect is larger for the realistic NFW model in the $\\Lambda$CDM universe than that for the self-similar model in the Einstein--de Sitter universe, assuming that they have the same $f_{\\rm gas}$. Our simulations suggest that for relaxed clusters with $M_{\\rm sh}\\sim 1.5 \\times 10^{15} M_{\\odot}$, the non-equipartition effect can account for only about 2\\%--3\\% of the missing thermal energy globally. For the most massive clusters, up to 4\\%--5\\% of the thermal energy beyond the equipartition value may be stored in the thermal energy of ions near the shock radius, but for clusters with $M_{\\rm sh} \\lesssim 5 \\times 10^{14} M_{\\odot}$, the non-equipartition effect is less than $1 \\%$. Thus, we argue that, at least for relaxed clusters, the non-equipartition effect alone can only account for some of the missing thermal energy problem, if any, for high-mass clusters but not for clusters with smaller masses. On the other hand, this suggests that hot gas may be missing due to other astrophysical processes not yet known, and the $f_{\\rm gas}$ in a real cluster should be lower than that we used in our numerical simulations. We have estimated that reducing $f_{\\rm gas}$ by $20\\%$ will enhance the local non-equipartition effect near the outer region by a few percent, but the integrated SZ bias, $Y_{\\rm non{\\text -}eq}/Y_{\\rm eq}$ is not affected by more than a percent. We emphasis here that even though non-equipartition effects may not affect the global energy budget significantly, the effect is still important locally in the outer regions ($\\sim R_{\\rm vir}$) of a cluster. Future X-ray and SZ observations may extend out to $R_{\\rm sh}$, and the effect of non-equipartition should be considered when studying cluster properties in those regions. We found that for our realistic NFW model in the $\\Lambda$CDM universe, $Y_{\\rm non{\\text -}eq}/Y_{\\rm eq}$ evolves with redshift, which is in contrast to the self-similar model in the Einstein--de Sitter universe. For our realistic NFW model in the $\\Lambda$CDM universe, $Y_{\\rm non{\\text -}eq}/Y_{\\rm eq}$ decreases as $z$ decreases. This is probably due to the decreasing rate of accretion onto clusters in the $\\Lambda$CDM universe during the period of cosmological acceleration, which results in a relatively longer timescale for the electron--ion equilibration inside a cluster compared to a cluster with the same mass in the Einstein--de Sitter universe. Though the magnitude of $Y_{\\rm non{\\text -}eq}/Y_{\\rm eq}$ is small for the range of cluster masses, even a percentage level deviation in the most massive clusters can be important for precision cosmology studies. Such a variation of $Y$ with $z$ would introduce an apparent evolution in $f_{\\rm gas}$, which would bias the cosmological studies using the $f_{\\rm gas}$ techniques \\citep{ARS08}. Recently, \\citet{RN09} have shown that the non-equipartition effect on $Y$ can be enhanced by major mergers up to $30\\%$, although for low Mach number mergers, the shock heating efficiency for electrons may be higher which can weaken the non-equipartition effect \\citep{GLR07, MV07}. The temporary boost due to mergers may have a significant effect on the estimation of cosmological parameters using clusters. We defer a detailed study of the effect on cosmology studies in a future paper. K.W. thanks Avi Loeb and Brian Mason for helpful discussions. Support for this work was provided by the National Aeronautics and Space Administration, through {\\it Chandra} Award Numbers TM7-8010X, GO7-8135X, GO8-9083X, and GO9-0035X, NASA $XMM-Newton$ grants NNX08AZ34G, NNX08AW83G, and NASA $Suzaku$ grant NNX08AI27G. We thank the referee for helpful comments. \\appendix" }, "0910/0910.0258_arXiv.txt": { "abstract": "We describe and test a new method for creating initial conditions for cosmological N-body dark matter simulations based on second-order Lagrangian perturbation theory (2lpt). The method can be applied to multi-mass particle distributions making it suitable for creating resimulation, or `zoom' initial conditions. By testing against an analytic solution we show that the method works well for a spherically symmetric perturbation with radial features ranging over more than three orders of magnitude in linear scale and eleven orders of magnitude in particle mass. We apply the method and repeat resimulations of the rapid formation of a high mass halo at redshift $\\sim50$ and the formation of a Milky-Way mass dark matter halo at redshift zero. In both cases we find that many properties of the final halo show a much smaller sensitivity to the start redshift with the 2lpt initial conditions, than simulations started from Zel'dovich initial conditions. For spherical overdense regions structure formation is erroneously delayed in simulations starting from Zel'dovich initial conditions, and we demonstrate for the case of the formation the high redshift halo that this delay can be accounted for using the spherical collapse model. In addition to being more accurate, 2lpt initial conditions allow simulations to start later, saving computer time. ", "introduction": "Computer simulations of cosmological structure formation have been crucial to understanding structure formation particularly in the non-linear regime. Early simulations e.g. \\cite{Aarseth79} used only around a thousand particles to model a large region of the Universe, while recent simulations of structure formation have modelled over a billion particles in just a single virialised object \\citep{Springel08,Stadel09}. Since the advent of the Cold Dark Matter (CDM) model \\citep{Peebles82,DEFW85}, most computational effort has been expended modelling structure formation in CDM universes. Early work in the 1980s focused on `cosmological' simulations where a representative region of the universe is modelled, suitable for studying large-scale structure \\citep{DEFW85}. The starting point for these CDM simulations, the initial conditions, are Gaussian random fields. The numerical techniques needed to create the initial conditions for cosmological simulations were developed in the 1980s and are described in \\cite{Efstathiou85}. As the algorithms for N-body simulations have improved, and the speed of computers increased exponentially with time, it became feasible in the 1990s to simulate the formation of single virialised halos in the CDM model with enough particles to be able to probe their internal structure \\citep[e.g.][]{NFW96, NFW97, Ghigna98, Moore99}. These more focused simulations required new methods for generating the initial conditions. The algorithm of \\cite{HoffmanRibak91} for setting up constrained Gaussian random fields was one method which could be applied to setting up initial conditions for a rare peak where a halo would be expected to form. The essence of this technique is that it allows selection of a region based on the properties of the linear density field. However, the objects we actually observe in the Universe are the end products of non-linear structure formation and it is desirable, if we want to understand how the structure we see formed, to be able to select objects on the basis of their final properties. This requirement led to an alternative method for producing initial conditions based first on selecting objects from a completed simulation e.g. \\cite{Katz93}, \\cite{Navarro94}. The initial conditions for the first or parent simulation, were created using the methods outlined by \\cite{Efstathiou85}. The density field is created out of a superposition of plane waves with random phases. Having selected an object at the final (or any intermediate) time from the parent simulation a fresh set of initial conditions with higher numerical resolution in the region of interest, which we will call `resimulation' initial conditions (also called `zoom' initial conditions) can be made. Particles of different masses are laid down to approximate a uniform mass distribution, with the smallest mass particles being concentrated in the region from which the object forms. The new initial conditions are made by recreating the the same plane waves as were present in parent simulation together with the addition of new shorter wavelength power. An alternative technique for creating resimulation initial conditions was devised by \\cite{Bertschinger01} based on earlier work by \\cite{Salmon96} and \\cite{Pen97} where a Gaussian random field with a particular power spectrum is created starting from a white noise field. Recently parallel code versions using this method has been developed to generate very large initial conditions \\citep{Prunet08,Stadel09}. A feature common to both these techniques, to date, is that the displacements and velocities are set using the Zel'dovich approximation \\citep{Zeldovich70}, where the displacements scale linearly with the linear growth factor. It has long been known that simulations starting from Zel'dovich initial conditions exhibit transients and that care must be taken choosing a sufficiently high start redshift so that these transients can decay to a negligible amplitude \\citep[e.g][]{Efstathiou85}. A study of the behaviour of transients using Lagrangian perturbation theory by \\cite{Scoccimarro98}, showed that for initial conditions based on second-order Lagrangian perturbation theory, the transients are both smaller and decay more rapidly than first-order, or Zel'dovich, initial conditions. \\cite{Scoccimarro98} gave a practical method for implementing second-order Lagrangian initial conditions. The method has been implemented in codes for creating cosmological initial conditions by \\cite{Sirko05} and \\cite{Crocce_etal06}. These codes are suitable for making initial conditions for cosmological simulations, targeted at large-scale structure, but they do not allow the creation of resimulation initial conditions, the focus of much current work on structure formation. In this paper we describe a new method for creating second-order Lagrangian initial conditions which can be used to make resimulation initial conditions. The paper is organised as follows: in Section~\\ref{SECTMOT} we introduce 2lpt theory initial conditions and motivate their advantages for studying structure in the non-linear regime by applying them to the spherical top-hat model; in Section~\\ref{MAKERESIMS} we describe how Zel'dovich resimulation initial conditions are made, and the new method for creating 2lpt initial conditions, and test the method against an analytic solution for a spherically symmetric perturbation; in Section~\\ref{REALAPPLICATIONS} we apply the method and analyse the formation of a dark matter halo at high and low redshift for varying starting redshifts for both Zel'dovich and 2lpt initial conditions; in Section~\\ref{SECTSUMMARY} we summarise and discuss the main results; in the Appendix we evaluate two alternate interpolation schemes which have been used in the process of making resimulation initial conditions. ", "conclusions": "} We have implemented a new method to make second order Lagrangian perturbation theory (2lpt) initial conditions which is well suited to creating resimulation initial conditions. We have tested the method for an analytic spherically symmetric perturbation with features ranging over a factor of 5000 in linear scale or eleven orders of magnitude in mass, and can reproduce the analytic solution to a fractional accuracy of better than one percent over this range of scales in radius measured from the symmetry centre. Applying the new method we have recreated the initial conditions for the formation of a high redshift halo from \\cite{Gao05} and the a Milky-way mass halo from \\cite{Springel08}. We find from studying the properties of the final halos, that the final conditions show much less sensitivity to the start redshift when using 2lpt initial conditions than when started from Zel'dovich initial conditions. For didactic purposes, we have calculated the effect of using Zel'dovich and 2lpt initial conditions for a spherical top-hat collapse. In this simple case, the epoch of collapse is delayed by a timing error which depends primarily on the number of expansions between the epoch of the initial conditions and the collapse time. This timing error, for fixed starting redshift, grows with collapse time for Zel'dovich initial conditions, but decreases with collapse time for 2lpt initial conditions. Applying the top-hat timing errors as a correction to the radius of shells containing fixed amounts of mass as a function of time has the effect of bringing the Zel'dovich and 2lpt resimulations of the high redshift \\cite{Gao05} halo into near coincidence. We find that for the Aq-A-5 halo the choice of start redshift for the Zel'dovich initial conditions in \\cite{Springel08} was sufficiently high at $z_s=127$ to make little difference to the properties of the halo at $z=0$ when compared to runs starting from 2lpt initial conditions. For lower starting redshifts ($z_s = 63, 31, 15$) the positions of substructures at $z=0$ do become sensitive to the precise start redshift. \\cite{Knebe09} have also looked at the properties of halos, for the mass range ($10^{10} - 10^{13}h^{-1}M_\\odot$) at redshift zero, for simulations starting from Zel'dovich and 2lpt initial conditions (generated using the code by \\cite{Crocce_etal06}). In their paper they simulated a cosmological volume which yielded many small halos and about ten halos with 150000 or more particles within the virial radius. In their study they looked at the distribution of halo properties including the spin, triaxiality and concentration. They also matched halos in a similar way to the way we have matched subhalos and looked at the ratios of masses, triaxialities, spins and concentrations of the matching halos. The authors concluded that any actual trends with start redshift or type of initial condition by redshift zero are certainly small (for start redshifts of 25 - 150). It is not possible to directly compare our results with theirs, but the trends in the halo properties we have observed in halo Aq-A-5 at redshift zero are indeed small, and not obviously in disagreement with their results for populations of halos. For many purposes, the differences which arise from using Zel'dovich or 2lpt initial conditions may be small when compared to the uncertainties which arise in the modelling of more complex physical processes. From this point of view the main advantage of using 2lpt initial conditions is that they allow the simulations to start later thus saving computer time. The issue of exactly what redshift one can start depends on the scientific problem, but the spherical collapse model, discussed in Subsection~\\ref{substh} gives a quantitative way to compare the start redshift for Zel'dovich or 2lpt initial conditions. However it remains the case that it is advisable to test the sensitivity of simulation results to the start redshift by direct simulation as for example in \\cite{Power03} for resimulations of galactic dark matter halos." }, "0910/0910.4198_arXiv.txt": { "abstract": "We report on the Camera Materials Test Chamber, a multi-vessel apparatus which analyzes the outgassing consequences of candidate materials for use in the vacuum cryostat of a new telescope camera. The system measures the outgassing products and rates of samples of materials at different temperatures, and collects films of outgassing products to measure the effects on light transmission in six optical bands. The design of the apparatus minimizes potential measurement errors introduced by background contamination. ", "introduction": "Introduction} The Large Synoptic Survey Telescope (LSST) is a planned facility which will repeatedly and deeply image large (10 square degree) sections of the night sky at optical and very near infrared wavelengths to an unprecidented depth, gathering important data to address outstanding questions in cosmolgy and astrophysics\\cite{LSSTover}. The etendue, or light gathering ability, of LSST, as measured by the primary mirror aperture size multiplied by the field of view, is more than an order of magnitude higher than any existing or planned optical facility. Such imaging requires a 3.2 billion pixel cryogenic CCD camera maintained in a clean vacuum environment. The cryostat of the LSST camera will measure 2.9 $m^{3}$, contain nearly 1000 kg of material, and have more than 40 $m^{2}$ of exposed surface area within, including the cold focal plane CCD detectors, two stages of supporting electronics, mechanical and thermal control structures, electrical cabling, electrical and fluid connectors and feed-thrus, and devices for internal metrology. All materials within the evacuated cryostat will remain there for years and be subject to thermal cycling. These materials must not outgass in a way that compromises the insulating vacuum nor change the net light transmittance to the focal plane between calibrations. There is scant data in the literature in regard to how the outgassing and deposition of commercial and other materials may affect light transmittance to optical surfaces in vacuum. Abromovici et al. limited the effect of elastomer outgassing on mirror reflectivities\\cite{ocon}. They use resonant cavities pumped by a laser, thus measuring at only one fequency, and do not cool the optical collecting surfaces (mirrors). Others have measured the outgassing and deposition rates of various matrials for many applications. However, given the almost endless variety of available materials, achieving a database relevant to very different applications is impossible. We have designed, contstructed, and commissioned an apparatus, the Camera Materials Test Chamber (CMTC), to quantiy the suitability of candidate materials for inclusion in the LSST camera cryostat. The CMTC could also be used to perform suitability tests for any similar vacuum application, and, given its ability to isolate functions and cold stages, can be used for a variety of precision vacuum measurements, such as the performance of getters. We report on the design and performance of the instrument, highlighting vacuum and thermal engineering considerations that may be of use to experimentalists. ", "conclusions": "" }, "0910/0910.4960_arXiv.txt": { "abstract": "For half a century, evidence has been growing that the formation of stars follows a universal distribution of stellar masses. In fact, no stellar population has been found showing a systematic deviation from the canonical initial mass function (IMF) found for example for the stars in the solar neighbourhood. The only exception may be the young stellar discs in the Galactic Centre, which have been argued to exhibit a top-heavy IMF. Here we discuss the question whether the extreme circumstances in the centre of the Milky Way may be the reason for a significant variation of the IMF. By means of stellar evolution models using different codes, we show that the observed luminosity in the central parsec is too high to be explained by a long-standing top-heavy IMF as suggested by other authors, considering the limited amount of mass inferred from stellar kinematics in this region. In contrast, continuous star formation over the Galaxy's lifetime following a canonical IMF results in a mass-to-light ratio and a total mass of stellar black holes (SBHs) consistent with the observations. Furthermore, these SBHs migrate towards the centre due to dynamical friction, turning the cusp of visible stars into a core as observed in the Galactic Centre. For the first time here we explain the luminosity and dynamical mass of the central cluster and both the presence and extent of the observed core, since the number of SBHs expected from a canonical IMF is just enough to make up for the missing luminous mass. We conclude that observations of the Galactic Centre are well consistent with continuous star formation following the canonical IMF and do not suggest a systematic variation as a result of the region's properties such as high density, metallicity, strong tidal field etc. If the young stellar discs prove to follow a top-heavy IMF, the circumstances that led to their formation must be very rare, since these have not affected most of the central cluster. ", "introduction": "In \\citeyear{s55}, Salpeter found that the initial mass distribution of field stars in the range $0.4\\simless M_{\\star}/\\msun \\simless 10$ is a power-law with exponent 2.35. Since then, a large number of publications have investigated the initial mass function (IMF) of stars and made clear that star formation in general follows the same empirical law, the canonical IMF \\citep[and references therein]{k01}. Due to its extreme conditions (mass density, velocity dispersion, tidal forces), the Galactic Centre provides a unique environment for testing the universality of the IMF. Star formation in the central region has thus been studied in detail, however no agreement has been reached on the nature of the IMF in either theory or observations: \\citet{mmt07} find a best fit of observations in the central parsec of our Galaxy with a model of constant star formation with a top-heavy IMF. On the other hand, \\citet{bse09} show that the old stellar cluster in the Galactic Centre very well resembles the bulge population. Observations of the young, massive Arches cluster in the central region of the Milky Way have long been interpreted as a prime example for top-heavy star formation \\citep[e.g.][]{fetal99,setal02,kfkn06,ksj07}. However, \\citet{esm09} have shown that a canonical IMF cannot be excluded for this cluster. \\citet{pau06} suggested a flat IMF for the young OB-stars observed in discs in the central parsec from the analysis of the K-band luminosity function. Based on more recent spectroscopic observations, \\citet{bmt+09} find strong evidence for this to be true. \\citet{br08} found from SPH simulations that the IMF of stars forming in fragmenting accretion discs strongly depends on the parameters of the underlying gas infall scenario. Unfortunately, theoretical IMF predictions have failed in the past to correctly describe the observations near the Galactic Centre \\citep{k08}. In this paper, we combine observational data with models of stellar evolution and dynamics to constrain the stellar mass function and star formation history in the Galactic Centre. It is organised as follows: In Section~\\ref{sec:ssemodels}, we analyse the properties of models of the Galactic Centre assuming different star formation histories, and compare them to the observations. Section ~\\ref{sec:massprof} describes the mass profile of the central parsec and the effect of mass segregation. We discuss the IMF of the young stellar discs around \\sgra\\ and appropriate formation scenarios in Section~\\ref{sec:discimf} and summarise in Section~\\ref{sec:conclusions}. ", "conclusions": "\\label{sec:conclusions} Various attempts have been made to study star formation in the Galactic Centre, but so far neither theory or simulations nor observations led to an agreement on the (initial) distribution of stellar masses. Here we have shown that theory and observations are consistent with star formation generally following a canonical IMF \\citep{k01} in the Galactic Centre, just as anywhere else in the Universe. Our main results can be summarised as follows: \\renewcommand{\\labelenumi}{\\arabic{enumi}.} \\begin{enumerate} \\item The mass-to-light ratio of the central parsec of the Milky Way is consistent with a constant or exponentially decreasing star formation rate following a canonical IMF. Models of constant star formation following an IMF with $\\alpha = 1.35$ are consistent with the observed luminosities but create $\\sim 10^5$ SBHs in the central parsec, ten times more than expected by other authors. Mass functions flatter than $\\alpha \\approx 1$ can be safely ruled out, since they cannot explain the observed number of bright stars and the diffuse light. \\item The core observed in the luminosity distribution with a radius of $r_{\\rm break} \\approx 0.2$\\,pc does not imply a core in the mass profile, but can be well explained by mass segregation as suggested by \\citet{fak06}, where dark remnants mass-dominate this region. Again, the observations are best explained by a canonical IMF, and are not compatible with star formation following a top-heavy IMF with $\\alpha \\le 1.35$, as this would create a core radius one order of magnitude larger. \\item Recent observations revealing a deficit of B-type stars in the young stellar discs suggest a top-heavy IMF for this population, which may be explained by tidal stripping of an infalling mass-segregated cluster, or unusual modes of star formation in a fragmenting accretion disc. However, these results do not allow conclusions on star formation in the Galactic Centre in general, for which we have no reason to assume it to be non-canonical. \\end{enumerate} While other authors generally predicted a flat IMF \\citep[e.g.][]{ksj07} or a higher low-mass cut-off \\citep[e.g.][]{m93,lb03,l_06} from state-of-the-art theoretical star formation models of the Galactic Centre, we find that observations suggest star formation follows a canonical IMF even under the extreme circumstances present in the central cluster. This universality of the IMF poses a major challenge to our understanding of star formation processes (see also \\citealp{k01,k08b}). Our results rely on the assumption that the stars observed within 1\\,pc from the Galactic Centre also formed there, suggesting star formation with a canonical IMF even under such exotic conditions. It is possible that some of the stars were brought in by massive star clusters which spiralled towards the SMBH through dynamical friction \\citep{pbmmhe06, fifm08}. However, this scenario is unlikely because a star cluster is stripped on its way towards the centre and loses mostly low-mass stars, since it would be in a mass segregated state soon after its formation. Therefore, the most likely scenario is a central cluster that formed over a Hubble time with a canonical IMF, where the very young stellar population of the stellar discs observed to have a very top-heavy IMF \\citep{bmt+09} constitutes a rare star formation event not typical for the bulk stellar population in the central cluster." }, "0910/0910.4367_arXiv.txt": { "abstract": "We report the discovery of radio afterglow emission from the gamma-ray burst GRB\\,090423, which exploded at a redshift of 8.3, making it the object with the highest known redshift in the Universe. By combining our radio measurements with existing X-ray and infrared observations, we estimate the kinetic energy of the afterglow, the geometry of the outflow and the density of the circumburst medium. Our best fit model is a quasi-spherical, high-energy explosion in a low, constant-density medium. \\event\\ had a similar energy release to the other well-studied high redshift GRB 050904 ($z=6.26$), but their circumburst densities differ by two orders of magnitude. We compare the properties of \\event\\ with a sample of GRBs at moderate redshifts. We find that the high energy and afterglow properties of \\event\\ are not sufficiently different from other GRBs to suggest a different kind of progenitor, such as a Population III star. However, we argue that it is not clear that the afterglow properties alone can provide convincing identification of Population III progenitors. We suggest that the millimeter and centimeter radio detections of \\event\\ at early times contained emission from a reverse shock component. This has important implications for the detection of high redshift GRBs by the next generation of radio facilities. ", "introduction": "\\label{sec:intro} Because of their extreme luminosities GRBs are detectable out to large distances by current missions, and due to their connection to core collapse SNe \\citep{wb06}, they could in principal reveal the stars that form from the first dark matter halos ($z\\sim$ 20--30) through to the epoch of reionization at $z=11\\pm 3$ and closer \\citep{lr00,cl00b,gmaz04,ioc07}. As bright continuum sources, GRB afterglows also make ideal backlights to probe the intergalactic medium as well as the interstellar medium in their host galaxies. Predicted to occur at redshifts beyond those where quasars are expected, they could be used to study both the reionization history and metal enrichment of the early universe \\citep{tkk+06}. The fraction of detectable GRBs that lie at high redshift ($z>6$) is, however, expected to be small ($<$10\\%; \\citealt{pcb+09,bl06}). Until recently there were only two GRBs with measured redshifts $z>6$; GRB\\,050904 \\citep{kka+06} and GRB\\,080913 \\citep{gkf+09} with $z=6.3$ and $z=6.7$, respectively. However, on April 23, 2009 the \\Swift\\ Burst Alert Telescope (BAT) discovered \\event\\ and the on-board X-ray Telescope (XRT) detected and localized a variable X-ray afterglow \\citep{tfl+09,sdc+09}. In ground-based follow-up observations no optical counterpart was found but a fading afterglow was detected by several groups at wavelengths longward of J band (1.2 $\\mu$m). Based on both broadband photometry and near infrared (NIR) spectroscopy, the sharp optical/NIR drop off was argued to be due to the Lyman-$\\alpha$ absorption in the intergalactic medium, consistent with a redshift with a best-fit value of $z=8.26^{+0.07}_{-0.08}$ \\citep{tfl+09}. The high redshift of \\event\\ makes it the most distant observed GRB, as well as the most distant object of any kind other than the Cosmic Microwave Background. This event occurred approximately 630 million years after the Big Bang, confirming that massive stellar formation occurred in the very early universe. In this paper we report the discovery of the radio afterglow from \\event\\ with the Very Large Array\\footnote{The Very Large Array is operated by the National Radio Astronomy Observatory, a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc.} (VLA). Broadband afterglow observations provide constraints on the explosion energetics, geometry, and immediate environs of the progenitor star. The afterglow has a predictable temporal and spectral evolution that depends on the kinetic energy and geometry of the shock, the density structure of the circumburst environment, and shock microphysical parameters which depend on the physics of particle acceleration and the circumburst magnetic field. To the degree that we can predict differences in the explosion and circumburst media between GRB progenitors at high and low redshifts, we can search for these different signatures in their afterglows. This has been the motivation for previous multi-wavelength modeling of the highest-$z$ afterglows \\citep{fck+06,gfm07}. In order to investigate the nature of the \\event\\ explosion, we combine our radio measurements with published X-ray and NIR observations, and apply a model of the blast wave evolution to fit the afterglow data (\\S\\ref{sec:results}). We compare the explosion energetics, circumburst density, and other derived characteristics to a sample of well-studied events, and discuss prospects for using afterglow measurements to investigate the nature of high-$z$ massive star progenitors. ", "conclusions": "\\label{sec:dis} \\event\\ is the highest-redshift object for which we have multi-wavelength observations, including good quality radio measurements. Below we address the following questions: based on its afterglow properties what can we learn about properties of the explosion and environs for this highest-redshift GRB? And, can we identify any differences between high and low redshift GRBs which indicate that they might arise from different progenitors? In particular, the initial generations of stars in the early universe are thought to be brighter, hotter and more massive ($>100\\, M_\\odot$) than stars today \\citep{hai08,byhm09}. Detecting these so-called Population III (Pop III) stars is one of the central observational challenges in modern cosmology, and the best prospect appears to be through observing their stellar death \\citep{hfw+03} via a supernovae (SNe) or gamma-ray burst explosion. It is worth asking what observational signatures could signal a Pop III GRB. Other than \\event , only one other $z > 6$ event, GRB\\,050904 ($z=6.26$), has high quality broadband afterglow measurements. In Fig.~\\ref{fig:comparison} we plot the best-fit parameters of these two GRBs along with a sample of well-studied lower redshift events from \\citet{pk01}. Both high redshift bursts stand out in terms of their large blast wave energy ($> 10^{52}$ erg). We know from samples of well-studied afterglows \\citep{fks+01,pk01,yhsf03}, that most have radiative and kinetic energies of order $\\sim 10^{51}$ erg. In the collapsar model the jet kinetic energy from a Pop III GRB could be 10--100 times larger than a Population II (Pop II) event \\citep{fwh01,hfw+03}. However, an energetic explosion does not appear to be an exclusive property of high-$z$ GRBs. There is a small but growing population of bursts with energy $> 10^{52}$ erg, termed 'hyper-energetic GRBs' \\citep{cfh+09}, which includes moderate-$z$ events like GRB\\,070125 \\citep{ccf+08} and GRB\\,050820A \\citep{ckh+06}. Another potentially useful diagnostic is the density structure in the immediate environs of the progenitor star. The radio data is a sensitive {\\it in situ} probe of the density because its emission samples the optically thick part of the synchrotron spectrum. The afterglows of \\event\\ and GRB\\,050904 are best fit by a constant density medium and not one that is shaped by stellar mass loss \\citep{cl99}. However, many afterglows at all redshifts are best fit by a constant density medium (e.g.~\\citealt{yhsf03}). The density obtained for GRB\\,050904 was the highest seen ($n\\approx84-680$ cm$^{-3}$) for any GRB to date, while \\event\\ with $n=0.9$ cm$^{-3}$ does not stand out (Fig.~\\ref{fig:comparison}), indicating these two high redshift bursts exploded in very different environments. A circumburst density of order unity is predicted for Pop III stars, since this density is limited by strong radiation pressure in the mini halo from which the star was formed \\citep{byh03}. This is not an unique property, since many local SNe explode in tenuous media, and so density constraints are not useful to signal Pop III explosions. For the other afterglow parameters ($p$, $\\epsilon_e$, $\\epsilon_B$ and $\\theta_j$) there are no published predictions for how they may differ between different progenitor models. Thus we turn to considering the prompt high-energy emission of \\event. \\citet{sdc+09} and \\citet{tfl+09} both noted that the high-energy properties of \\event\\ (fluence, luminosity, duration, radiative energy) are not substantially different from those moderate-$z$ GRBs. We are not aware of any quantitative predictions based on these observed properties that would discriminate between Pop\\,II and Pop\\,III progenitors. For example, apart from the effect of ($1+z$) time dilation, there is no reason to expect high-$z$ GRBs to have significantly longer intrinsic durations. Collapsar models which form black holes promptly through accretion onto the proto-neutron star (Type I) or via direct massive ($>260$ M$_\\odot$) black hole formation (Type III) have durations set by jet propagation and disk viscosity timescales, respectively \\citep{mwh01,fwh01}, which are $\\sim 10$ s in the rest frame (with large uncertainties; \\citealt{fwh01}). Metallicity can also be an important discriminant. There is a critical metallicity ($Z>10^{-3.5}\\,Z_\\odot$) below which high-mass Pop III stars dominate \\citep{byhm09,bl06}. The contribution of Pop III stars to the co-moving star formation rate is expected to peak around $z=15$ but their redshift distribution exhibits a considerable spread to $z\\sim$7. Thus we might find high-redshift GRBs with Pop III progenitors in ``pockets'' of low metallicity. \\citet{sdc+09} argue for a lower bound of $Z>0.04\\,Z_\\odot$ based on their detection of excess soft X-ray absorption by metals along the line of sight, in comparison to the Milky Way column density predicted from \\ion{H}{1} (21\\,cm) measurements. We do not consider this a robust measurement as it is sensitive to a range of unaccounted-for systematic effects, including: spectral variability; spectral curvature; low-amplitude X-ray flares; and the presence of intervening (cosmological) absorption systems along the line of sight. Summarizing the above discussion, we do not find that the individual properties of \\event\\ are sufficiently dissimilar to other GRBs to warrant identifying it as anything other than a normal GRB. We lack robust predictions of well-defined afterglow signatures that could allow us to unambiguously identify a Pop III progenitor star from its afterglow properties alone. Significantly larger numbers of GRBs at high redshift with well-sampled afterglow light curves, high-resolution spectra, and host galaxy detections are needed to determine if high redshift GRB progenitors differ in a statistical sense from those at low redshift. We note that, like GRB\\,050904, the \\event\\ afterglow indicates the signature of reverse shock (RS) emission in the radio, as seen in the VLA and PdBI data. \\citep{ioc07} have studied the expected RS emission at high redshift, and they find that the effects of time dilation almost compensate for frequency redshift , resulting in a near-constant observed peak frequency in the mm band ($\\nu \\sim 200$~GHz) at a few hours post-event, and a flux at this frequency that is almost independent of redshift. Further, the mm band does not suffer significantly either from extinction (in contrast to the optical) or scintillation (in contrast to the radio). Therefore, detection of mm flux at a few hours post event should be a good method of indicating a high redshift explosion. ALMA, with its high sensitivity ($\\sim$75 $\\mu$Jy in 4 min), will be a potential tool for selecting potential high-$z$ bursts that would be high priority for intense followup across the spectrum. This will hopefully greatly increase the rate at which high-$z$ events are identified. Finally, our data does not rule out a late jet break at $t_j>45$ d, which, as discussed above, makes the total explosion energy uncertain. Extremely sensitive VLA observations would be required to distinguish between the isotropic versus jet model. However, in 2010 with an order of magnitude enhanced sensitivity the EVLA will be the perfect instrument for such studies. For a 2\\,hr integration in 8 GHz band, the EVLA can reach sensitivity up to 2.3 $\\mu$Jy which will be able to detect the \\event\\ for 2 years or 6 months if the burst is isotropic or jet-like, respectively. EVLA will thus be able detect fainter events and follow events like GRB\\,050904 and \\event\\ for a longer duration, therefore obtaining better density measurements, better estimates of outflow geometry and the total kinetic energy." }, "0910/0910.3771_arXiv.txt": { "abstract": "Spectral energy distributions (SEDs) of the central few tens of parsec region of some of the nearest, most well studied, active galactic nuclei (AGN) are presented. These genuine AGN-core SEDs, mostly from Seyfert galaxies, are characterised by two main features: an IR bump with the maximum in the 2 -10 $\\mu$m range, and an increasing X-ray spectrum with frequency in the 1 to $\\sim$ 200 keV region. These dominant features are common to Seyfert type 1 and 2 objects alike. In detail, type 1 AGNs are clearly distinguished from type 2s by their high spatial resolution SEDs: type 2 AGN exhibit a sharp drop shortward of 2 $\\mu$m, with the optical to UV region being fully absorbed; type 1s show instead a gentle 2 $\\mu$m drop ensued by a secondary, partially-absorbed optical to UV emission bump. On the assumption that the bulk of optical to UV photons generated in these AGN are reprocessed by dust and re-emitted in the IR in an isotropic manner, the IR bump luminosity represents $\\gtrsim 70\\%$ of the total energy output in these objects, the second energetically important contribution are the high energies above 20 keV. \\\\ Galaxies selected by their warm infrared colours, i.e. presenting a relatively-flat flux distribution in the 12 to 60 $\\mu$m range have often being classified as active galactic nuclei (AGN). The results from these high spatial resolution SEDs question this criterion as a general rule. It is found that the intrinsic shape of the infrared spectral energy distribution of an AGN and inferred bolometric luminosity largely depart from those derived from large aperture data. AGN luminosities can be overestimated by up to two orders of magnitude if relying on IR satellite data. We find these differences to be critical for AGN luminosities below or about $10^{44}$ erg~ s$^{-1}$. Above this limit, AGNs tend to dominate the light of their host galaxy regardless of the integration aperture size used. Although the number of objects presented in this work is small, we tentatively mark this luminosity as a threshold to identify galaxy-light- vs AGN- dominated objects. \\\\ ", "introduction": "} The study of the spectral energy distributions (SED) over the widest possible spectral range is an optimal way to characterise the properties of galaxies in general. Covering the widest spectral range is the key to differentiate physical phenomena which dominate at specific spectral ranges: e.g. dust emission in the IR, stellar emission in the optical to UV, non-thermal processes in the X-rays and radio, and to interrelate them as most of these phenomena involve radiation reprocessing from a spectral range into another. The availability of the SED of a galaxy allows us to determine basic parameters such as its bolometric luminosity (e.g. Elvis et al. 1994; Sanders \\& Mirabel 1996; Vasudevan \\& Fabian 2007), and via modelling of the SED, its star formation level, mass and age (e.g. Bruzual \\& Charlot 2003; Rowan-Robinson et al. 2005; Dale et al. 2007). The construction of bona-fide SEDs is not easy as it involves data acquisition from different ranges of the electromagnetic spectrum using very different telescope infrastructure. That already introduces a further complication as the achieved spatial resolution, and with it the aperture size used, vary with the spectral range. SEDs based on the integration of the overall galaxy light may be very different from those extracted from only a specific region, for example the nuclear region. In this specific case, the aperture size matters a lot, as different light sources, such as circumnuclear star formation, the active nucleus, and the subjacent galaxy light, coexist on small spatial scales and may contribute to the total nuclear output with comparable energies (e.g. Genzel et al. 1998; Reunanen et al. 2009). In the specific case of SEDs of AGN, it is often assumed that the AGN light dominates the integrated light of the galaxy at almost any spectral range and for almost any aperture. This assumption becomes mandatory at certain spectral ranges, such as the high energies, the extreme UV or the mid-to-far-IR, because of the spatial resolution limitations imposed by the available instrumentation, which currently lies in the several arcsecs to arcminutes at these wavelengths. In the mid- to far- IR in particular, the available data, mostly from IR satellites, are limited to spatial resolutions of a few arcsecs at best. Thus, the associated SEDs include the contribution of the host galaxy, star forming regions, dust emission and the AGN, with the first two components being measured over different spatial scales in the galaxy depending on the object distance and the spatial resolution achieved at a given IR wavelength. In spectral ranges where high spatial resolution is readily available, the importance, if not dominance, of circumnuclear star formation relative to that of the AGN has become clear in the UV to optical range (e.g. Munoz-Marin et al. 1997), or in the near-IR (Genzel et al. 1998). In the radio regime, the comparison of low- and high- spatial-resolution maps shows the importance of the diffuse circumnuclear emission and the emission from the jet components with respect to that of the core itself (e.g Roy et al., 1994; Elmouttie et al, 1998; Gallimore et al. 2004; Val, Shastri \\& Gabuzda, 2004). Even with low resolution data a major concern shared by most works is the relevance of the host galaxy contribution to the nuclear integrated emission from the UV to optical to IR. To overcome these mixing effects introduced by poor spatial resolution, different strategies or assumptions have been followed by the community. In quasars, by their own nature, the dominance of the AGN light over the integrated galaxy light at almost any wavelength is assumed; conversely, in lower luminosity AGN, the contribution of different components is assessed via modelling of the integrated light (Edelson \\& Malkan 1986; Ward et al. 1987; Sanders et al. 1988; Elvis et al. 1994; Buchanan et al. 2006 among others). In this paper, we attempt to provide a best estimate of the AGN light contribution on very nearby AGN by using very high spatial resolution data over a wide range of the electromagnetic spectrum. Accordingly, SEDs of the central few hundred parsec region of some of the nearest and brightest AGN are compiled. The work is motivated by the current possibility to obtain subarcsec resolution data in the near-to-mid-IR of bright AGN, and thus at comparable resolutions to those available with radio interferometry and the HST in the optical to UV wavelength range. This is possible thanks to the use of adaptive optics in the near-IR, the diffraction limit resolutions provided by 8 -10 m telescopes in the mid-IR as well as interferometry in the mid-IR. The selection of targets is driven by the requirements imposed by the use of adaptive optics in the near-IR, which limits the observations to the availability of having bright point-like targets with magnitudes V$~< $15 mag in the field, and the current flux detection limits in mid-IR ground-based observations. AGN in the near universe are sufficiently bright to satisfy those criteria. The near- to mid- IR high resolution data used in this work come mostly from the ESO Very Large Telescope (VLT), hence this study relies on Southern targets, all well known objects, mostly Seyfert galaxies: Centaurus A, NGC 1068, Circinus, NGC 1097, NGC 5506, NGC 7582, NGC 3783, NGC 1566 and NGC 7469. For comparison purposes, the SED of the quasar 3C 273 is also included. The compiled SEDs make use of the highest spatial resolution data available with current instrumentation across the electromagnetic spectrum. The main sources of data include: VLA-A array and ATCA data in radio, VLT diffraction-limited images and VLTI interferometry in the mid-infrared (mid-IR), VLT adaptive-optics images in the near-infrared, and \\textit{HST} imaging and spectra in the optical-ultraviolet. Although X-rays and $\\gamma$-rays do not provide such a fine resolution, information when available for these galaxies are also included in the SEDs on the assumption that above 10 keV or so we are sampling the AGN core region. Most of the data used comes from the Chandra and INTEGRAL telescopes. The novelty in the analysis is the spatial resolutions achieved in the infrared (IR), with typical full-width at half-maximum (FWHM) $\\lesssim $ 0.2 arcsec in the 1--5 $\\mu$m, $< $0.5 arcsec in the 11 -- 20 $\\mu$m. The availability of IR images at these spatial resolutions allow us to pinpoint the true spatial location of the AGN -- which happens not to have an optical counterpart in most of the type 2 galaxies studied -- and extract its luminosity within aperture diameters of a few tens of parsec. The new compiled SEDs are presented in sect. 3. Some major differences but also similarities between the SEDs of type 1 and type 2 AGN arise at these resolutions. These are presented and discussed in sections 4 and 5. The SEDs and the inferred nuclear luminosities are further compared with those extracted in the mid-to-far IR from large aperture data from IR satellites, and the differences discussed in sect. 6. Throughout this paper, $H_0 = 70$ km s$^{-1}$ Mpc$^{-1}$ is used. The central wavelength of the near-IR broad band filters used are: $I$-band (0.8 $\\mu$m), $J$-band (1.26 $\\mu$m), $H$-band (1.66 $\\mu$m), $K$-band (2.18 $\\mu$m), $L$-band (3.80 $\\mu$m) and $M$-band (4.78 $\\mu$m). \\\\ ", "conclusions": "Sub-arcsec resolution data spanning the UV, optical, IR and radio have been used to construct spectral energy distributions of the central, several tens of parsec, region of some of the nearest and brightest active galactic nuclei. Most of these objects are Seyfert galaxies. \\\\ These high spatial resolution SEDs differ largely from those derived from large aperture data, in particular in the IR: the shape of the SED is different and the true AGN luminosity can get overestimated by orders of magnitude if based on IR satellite data. These differences appear to be critical for AGN luminosities below $10^{44} erg~s^{-1}$ in which case large aperture data sample in full the host galaxy light. Above that limit we find cases among these nearby Seyfert galaxies where the AGN behaves as the most powerful quasars, dominating the host galaxy light regardless of the integration aperture-size used. \\\\ The high spatial resolution SED of these nearby AGNs are all characterised by two major features in their power distribution: an IR bump with maximum in the 2 -10 $\\mu$m range, and an increasing trend in X-ray power with frequency in the 1 to $\\sim$ 200 keV region, i.e. up to the hardest energy that was possible to sample. These dominant features are common to Seyfert type 1 and 2 objects alike. \\\\ The major difference between type 1 and 2 in these SEDs arises shortward of 2 $\\mu$m. Type 2s are characterised by a sharp fall-off shortward of this wavelength, with no optical counterpart to the IR nucleus being detected beyond 1 to 0.8 $\\mu$m. Type 1s show also a drop shortward of 2 $\\mu$m but this is more gentle - the spectrum is flatter - and recovers at about 1 $\\mu$m to give rise to the characteristic blue-bump feature seen in quasars. The flattening of the spectrum shortward of 2 $\\mu$m is also an expected feature of type 1 AGNs. Interpreting the IR bump as AGN reprocessed emission by the nuclear dust, in type 1s the nearest to the centre hotter dust can be directly seen, hence the flattening of their spectrum, whereas in type 2s this hot dust is still fully obscured. \\\\ Longward of 2 $\\mu$m, all the AGN types show very similar SEDs, the bulk of the IR emission starts from this wavelength on and the shape of the IR bump is very similar in all the AGNs. This is compatible with an equivalent black-body temperature for the bulk of the dust in the 200 - 400 K range in average. Although the current shape of the IR bump is limited by the availability of high angular resolution data beyond 20 $\\mu$m for most objects, due to the small region sampled in these SEDs, of just a few parsecs in some galaxies, a major contribution from colder dust that will modify the IR bump is not expected. \\\\ It can thus be concluded that at the scales of a few tens of parsec from the central engine, the bulk of the IR emission in either AGN type can be reconciled with pure dust emission. It follows that further contributions from a non-thermal synchrotron component and/or a thermal free-free emission linked to cooling of ionised gas are insufficient to overcome that of dust at these physical scales. The detailed modelling of NGC 1068's SED - this being one of the most complete we have compiled - in which these three contributions -- synchrotron, free-free and a dust torus components - are taking into account illustrates that premise, that is, the dominance of dust emission in the IR, even at the parsec-scale resolution achieved for this object in the mid-IR with interferometry (Hoenig et al. 2008). Only the two more extreme objects in this analysis, Cen A, on the low luminosity rank, and 3C 273, on the highest, present a SED that is not dominated by dust but by a synchrotron component. We tend to believe that is due to a much reduced dust content in these nuclei. \\\\ Over the nine orders of magnitude in frequency covered by these SEDs, the power stored in the IR bump is by far the most energetic fraction of the total energy budget measured in these objects. Evaluating this total budget as the sum of the IR and hard X-ray -- above 20 keV -- luminosities, the IR part accounts for more than 70\\% of the this total in seven out of the ten AGN studied. In the three exceptions, the IR fraction reduces to $<\\sim 30\\% $ (3C 273 and NGC 1566), $< \\sim 60\\%$ in Cen A. Even if accounting for variability in the X-rays, by a factor 2 to 3 in average, the IR emission remains in all cases dominant over, or as important as, in the last three cases, the X-ray emission. If comparing with the observed blue bump luminosity in the type 1 nuclei, this represents less than 15\\% of the IR emission. Putting all together, the IR bump energy from these high spatial resolutions SEDs may represent the tightest measurement of the accretion luminosity in these Seyfert AGN. \\\\ The average high spatial resolution SED of the type 2 and of the type 1 nuclei analysed in this work, and presented in Fig. 4, can be retrieved from http://www.iac.es/project/parsec/main/seyfert-SED-template. \\\\ This work was initiated and largely completed during the stay of K. Tristram, N. Neumayer and A. Prieto at the Max-Planck Institut fuer Astronomie in Heidelberg." }, "0910/0910.1774_arXiv.txt": { "abstract": "We present \\textit{B} and \\textit{R} band spectroastrometry of a sample of 45 Herbig Ae/Be stars in order to study their binary properties. All but one of the targets known to be binary systems with a separation of $\\rm \\sim0.1-2.0$~arcsec are detected by a distinctive spectroastrometric signature. Some objects in the sample exhibit spectroastrometric features that do not appear attributable to a binary system. We find that these may be due to light reflected from dusty halos or material entrained in winds. We present 8 new binary detections and 4 detections of an unknown component in previously discovered binary systems. The data confirm previous reports that Herbig Ae/Be stars have a high binary fraction, $\\rm{74\\pm6}$~per cent in the sample presented here. We use a spectroastrometric deconvolution technique to separate the spatially unresolved binary spectra into the individual constituent spectra. The separated spectra allow us to ascertain the spectral type of the individual binary components, which in turn allows the mass ratio of these systems to be determined. In addition, we appraise the method used and the effects of contaminant sources of flux. We find that the distribution of system mass ratios is inconsistent with random pairing from the Initial Mass Function, and that this appears robust despite a detection bias. Instead, the mass ratio distribution is broadly consistent with the scenario of binary formation via disk fragmentation. ", "introduction": "Our understanding of the formation and early evolution of massive stars ($\\rm M_* \\ga 8M_{\\odot}$) is much less complete than in the case of low mass stars. The scenario of low mass star formation has been relatively well studied, and a broadly consistent observational and theoretical picture has now emerged. The various phases of low mass star formation include: cloud collapse, proto-stellar creation and a subsequent contraction of Pre Main Sequence (PMS) objects towards the Zero Age Main Sequence (ZAMS). This later stage, the T Tauri phase, is easy to observe and therefore relatively well understood \\citep{Bouvier2007}. In the case of more massive stars the situation is much less clear. Such stars do not experience an optically visible PMS phase, evolve on a much more rapid timescale, and are considerably more luminous than low mass stars. Early studies on the effects of radiation pressure and the considerable ionising output of massive young stars prompted speculation that massive star formation might proceed in a different manner to that of low mass stars \\citep{Larson1971, Kahn1974}. For example, it has been suggested that the most massive stars form via stellar mergers or competitive accretion \\citep{JBally2005}. \\smallskip However, recent work, on both the observational and theoretical front, suggests that massive star formation may not be dissimilar to low mass star formation. As an example of observational results, \\citet{Pateletal2005} report the detection of a massive disk around a 15$\\rm M_{\\odot}$ protostar, indicating that massive stars may form via monolithic accretion. On the theoretical front, recent work indicates accretion onto a massive protostar is not impeded by radiation pressure \\citep{YorkeandSonnhalter2002,Turner2007,Krumholz2009}. However, while significant progress has been made, there remain many unaddressed questions related to the formation and evolution of massive stars \\citep{ZinneckerandYorke2007}. As observations of massive young stars are challenging, the full extent of the differences and similarities between low and high mass star formation are still unknown. \\smallskip Between the two extremes of mass lie the Herbig Ae/Be (HAe/Be) stars \\citep{Herbig1960}. These stars represent the most massive of objects to experience an optically visible PMS evolutionary phase. Therefore, HAe/Be stars offer an opportunity to study the early evolution of stars more massive than the sun. Spectropolarimetry indicates that Herbig Ae stars may undergo a PMS phase similar to that of the T Tauri stars, while Herbig Be stars may evolve via disk accretion, rather than magnetospheric accretion \\citep{JSVink2002,JSVink2005a,JCMottram2007}. Therefore, it appears that a transition in formation mechanisms occurs across the HAe/Be mass boundary \\citep{JCMottram2007}. However, the critical mass has not yet been established. \\smallskip To examine the similarities and differences between low mass T Tauri stars, HAe/Be stars and the optically invisible Massive Young Stellar Objects (MYSOs), study of the circumstellar environment at small angular scales is required. This is not trivial, requiring observations with high angular resolution \\citep{Mannings1997,Fuente2006,Grady2007,Kraus2008}. Despite the progress in the field, a full understanding of HAe/Be stars is hampered by the small sample sizes involved. By way of contrast, \\citet{DB2006} utilised spectroastrometry to study a large sample of HAe/Be stars with milli-arcsecond (mas) precision. Despite this resolution \\citet{DB2006} did not detect any accretion disks around HAe/Be stars. However, they did find that the majority, $\\rm{68\\pm11}$~per cent, of HAe/Be stars reside in relatively wide (probably a few-hundred~au, see Section \\ref{sep}) binary systems. \\smallskip The binary fraction reported by \\citet{DB2006} is greater than that of T Tauri stars at similarly wide separations, which in turn is greater than that of Main Sequence G-dwarfs at the same separations \\citep{DuquennoyandMayor1991,Reipurth1993,Ghez1993}. Indeed, this high binary fraction is approaching that of more massive stars \\citep{Preibisch1999}. However, little is known about the properties of such binary systems. The properties of the binary components and configurations of such systems are of interest as they can constrain the binary formation mechanism. The seminal study to date is that by \\citet{BC2001}, who used Adaptive Optics assisted observations to construct Spectral Energy Distributions (SEDs) for each component in a number of HAe/Be binary systems. The drawback of SED fitting is that PMS stars, as young stars, are inevitably associated with dusty, obscured environments. Therefore, the brightness ratio of a binary determined by SED fitting can occasionally be ambiguous. However, very few HAe/Be binary systems have been studied with spatially resolved spectroscopy, and thus far such studies have been conducted with seeing limited resolution \\citep{ACarmona2007,Hubrig2007}. \\smallskip The position angles of HAe/Be binary systems seem to be preferentially aligned with the spectropolarimetrically detected circumprimary disks \\citep{DB2006}. This already places constraints on the formation modes of these stars, in that it seems the systems formed via fragmentation of a molecular core or disk. This had already been suggested for lower mass binaries \\citep{SWolf2001,Kroupa2001}, but little is known about the formation mechanisms of more massive stars. This paper describes a spectroastrometric follow-up of the work of \\citet{DB2006} with dedicated observations to study both components of binary systems. The objective is to determine the properties of these binary systems and thus place stronger, more quantitative, constraints on the formation of stars of intermediate mass. We do this by determining the mass ratio of these binary systems. This is done using a spectroastrometric technique to disentangle the constituent spectra of unresolved binary systems, allowing the spectral type, and hence mass, of each component to be determined. Spectroastrometry itself is a relatively simple technique that extracts the spatial information present in conventional longslit spectra. Crucially, spectroastrometry can probe changes in flux distributions with a typical precision of a mas or less \\citep{JBailey1998a}, which is required to study unresolved binary systems. Typically the minimum separation probed is of the order 100~mas, as the signature of a binary system is dependant upon the system brightness ratio and separation. \\smallskip This paper is structured as follows: in Section \\ref{obsanddatred} we present our sample selection, observation method and data reduction procedures. In Section \\ref{results_spec_ast} we discuss the spectroastrometric signatures observed. In Section \\ref{spec_split_app} we present the method of splitting unresolved binary spectra and in Section \\ref{results_spec} we review the results of separating binary spectra into their constituent spectra. In Section \\ref{discussion} we discuss our results. Finally, we conclude this paper in Section \\ref{conclusions} by summarising the salient points raised. \\section[]{Observations and data reduction} \\label{obsanddatred} \\subsection{Observations} The data presented consist of long-slit spectra in the \\textit{B} band ($\\rm 4200-5000 \\rm{\\AA}$) and/or the \\textit{R} band ($\\rm 6200-7000 \\rm{\\AA}$) of 45 HAe/Be stars, and 2 emission line objects which are possible HAe/Be stars. The objects were chosen from the catalogs of \\citet{Theetal1994},\\,\\citet{Vieira2003}\\,\\&\\,\\citet{JHernandez2004}, and were selected to be reasonably bright (\\textit{V} $\\rm \\leq $ 12-13). Some objects previously observed by \\citet{DB2006} were observed to provide a consistency check on the spectroastrometric signatures. Given the small population of HAe/Be stars, the objects observed constitute a representative sample of HAe/Be stars, albeit brightness limited. \\smallskip The data were obtained using the 4.2m William Herschel Telescope (WHT) and the 2.5m Isaac Newton Telescope (INT). At the WHT, data were obtained on the 6th \\& 7th of October 2006, using the Intermediate Dispersion Spectrograph and Imaging System (ISIS) spectrograph. Spectra of 20 objects were taken simultaneously in the \\textit{B} and \\textit{R} bands using the dichroic slide of ISIS. In most cases a slit $\\rm{5}$~arcsec wide was used to ensure all the light from a given binary system entered the slit, even in poor seeing. This allows us to study the individual binary components, unlike \\citet{DB2006}, who used a slit of 1~arcsec. The R1200B and R1200R gratings were used and the resulting spectral resolving power was found to be $\\rm \\sim 3500$, corresponding to $\\rm 85\\, km\\,s^{-1}$. The angular pixel size was $\\rm{0.20}$ and $\\rm{0.22}$~arcsec in the \\textit{B} band and \\textit{R} band respectively, which means that the spatial profile of the longslit spectra was well sampled (average FWHM 1.9~arcsec). At the INT data were obtained using the 235mm camera and the Intermediate Dispersion Spectrograph (IDS). Observing was conducted from the 27th of December 2008 to the 3rd of January 2009. The spectra of 32 objects were obtained, despite adverse weather conditions preventing observing for the better part of three nights. As at the WHT the slit width was generally $\\rm5$~arcsec. The R1200R and R1200B gratings were used and the resulting spectral resolution was found to be $\\rm \\sim 3800 $, or $\\rm80 \\,km\\,s^{-1}$. The angular size of the pixels was $\\rm{0.4}$~arcsec, which fully sampled the average spatial profile of the spectra (1.8 arcsec). \\smallskip Multiple spectra were taken at four position angles (PA) on the sky. The PAs selected always comprised of two perpendicular sets of two anti-parallel angles, e.g. $\\rm 0\\degr$, $90^{\\circ}$, $180^{\\circ}$ and $270^{\\circ}$. Dispersion calibration arcs were made using CuNe and CuAr lamps. Table \\ref{logofobs} presents a summary of the observations. \\begin{center} \\begin{table*} \\begin{minipage}{\\textwidth} \\caption{\\label{logofobs}\\small{Log of the observations, column 1 lists the objects observed, column 2 denotes the spectral type of the objects, column 3 lists the \\textit{V} band magnitudes of the sample, and column 4 designates which telescope the object in question was observed with. Columns 5 and 6 list the average seeing conditions, columns 7 and 8 list the total exposure times and column 9 denotes the slit width used. Column 10 lists the total Signal to Noise Ratios, and finally, column 11 presents the date(s) each object was observed. Information on the objects is taken from SIMBAD (simbad.u-strasbg.fr) unless otherwise stated.}} \\begin{center} \\begin{tabular}{p{1.725cm} p{1.25cm} p{0.7cm} p{0.95cm} p{0.90cm} p{0.90cm} p{0.550cm} p{0.550cm} p{1.1cm} p {1.35cm} p{3.15cm}} \\hline Object & Spec type & \\hspace*{2mm}\\textit{V} &Telescope & $\\rm \\overline{FWHM}^a$ & $\\rm \\overline{FWHM}^b$& $\\rm{t_{blue}}$ &$\\rm{t_{red}}$ & Slit & SNR & Date\\\\ & & & &(\\arcsec)& (\\arcsec) & (s) & (s) & (\\arcsec) \\\\ \\hline \\hline VX Cas & $\\rm{A0e}$ & 11.3 & WHT & 1.3 &1.2 & 4800 & 4800 & 5.0&$\\rm{600_B}$,$\\rm{570_R}$ &07/10/2006 \\\\ VX Cas & $\\rm{A0e}$ & 11.3 & INT & 1.1 & 1.7 & 2800 & 3600 & $\\rm 2.5_{\\rm B},5.0_{\\rm R}$ & $\\rm{370_B}$,$\\rm{370_R}$&28/12/2008,31/12/2008\\\\ V594 Cas & Be & 10.6 & INT &1.3 & -- & 3200 &-- &5.0 &610 &01/01/2009\\\\ V1185 Tau & A1 & 10.7 & INT& 1.7 & --& 3200 &-- & 5.0 & 430&03/01/2009 \\\\ IP Per & $\\rm{A3}$ & 10.3 & INT & 1.2 & 1.4 &2000 & 2400 &$\\rm 2.5_{\\rm B},5.0_{\\rm R}$ & $\\rm{110_B}$,$\\rm{320_R}$&28/12/2008,31/12/2008\\\\ AB Aur & A0Vpe & 7.1 & WHT & 1.9 & 1.9 & 330 & 320 & 5.0 &$\\rm{110_B}$,$\\rm{650_R}$ &06/10/2006\\\\ MWC 480 & $\\rm{A3pshe}$& 7.7 & WHT & 2.0 & 2.1 & 960 & 640 & 5.0 & $\\rm{1100_B}$,$\\rm{800_R}$&06/10/2006\\\\ UX Ori & $\\rm{A3e}$ & 9.6 & WHT & 2.4 &2.4 & 3600 & 3600 & 5.0 & $\\rm{1200_B}$,$\\rm{940_R}$&06/10/2006\\\\ V1012 Ori & $\\rm{Be^c}$ &12.1 &INT &2.1& -- & 4800 & --& 5.0 &150 &02/01/2009\\\\ V1366 Ori & A0e& 9.8& INT& 1.3 &-- & 2400 & --& 3.0 &570 &31/12/2008 \\\\ V346 Ori & $\\rm{A5III}$&10.1 & INT & 1.5 & -- & 3600 &-- & 5.0 & 200&01/01/2009\\\\ HD 35929 & A5 & 8.1 & WHT & 1.9 & 1.7 & 2060 & 1470 & 5.0 & $\\rm{40_B}$,$\\rm{900_R}$&07/10/2006\\\\ V380 Ori & A0& 10.7 & INT & 1.5 &1.6 & 3600 & 2940 & $\\rm 3.0_{\\rm B},5.0_{\\rm R}$& $\\rm{100_B}$,$\\rm{200_R}$&28/12/2008,31/12/2008\\\\ MWC 758 & A3e & 8.3 & WHT & 1.4 & 1.3&1080 & 960& 5.0 &$\\rm{50_B}$,$\\rm{660_R}$ &07/10/2006\\\\ HK Ori & A4pev &11.9 & INT & 2.4 & -- & 4800 & --& 5.0 & 200&02/01/2009\\\\ HD 244604 & A3 & 9.4 & WHT & 1.7 & 1.6 & 3180 & 3660 & 5.0 &$\\rm{100_B}$,$\\rm{720_R}$ & 07/10/2006\\\\ V1271 Ori & A5 & 10.0 & INT &1.6&-- & 2460 & --& 5.0 & 410& 01/01/2009\\\\ T Ori & A3 & 9.5 & INT &1.9 & -- & 3200& -- & 5.0 & 300 &03/01/2009 \\\\ V586 Ori & $\\rm{A2V}$ &9.8 & INT & 3.3 & -- & 2940 & --& 5.0 & 650&02/01/2009\\\\ HD 37357 & A0e &8.8 & INT & 1.4 & 1.4 &2060 & 1470& $\\rm 3.0_{\\rm B},5.0_{\\rm R}$ & $\\rm{200_B}$,$\\rm{350_R}$ & 28/12/2008,31/12/2008\\\\ V1788 Ori & B9Ve & 9.9 & INT& 1.7 &-- & 1350 &-- & 5.0 & 450 &01/01/2009\\\\ HD 245906 & B9IV & 10.7 & INT & 1.8 & -- &2800 & --& 5.0 &100 &03/01/2009\\\\ RR Tau & A2II-IIIe & 10.9& INT & 1.7 & -- &2800 &-- & 5.0 & 250&03/01/2009 \\\\ V350 Ori & A0e & 10.4(\\textit{B})& INT & 1.9 & -- &4800 &-- & 5.0 & 130& 03/01/2009 \\\\ MWC 120 & A0 & 7.9 & WHT & 2.1 & 1.9 & 480 & 480 &5.0 &$\\rm{1500_B}$,$\\rm{690_R}$ & 06/10/2006\\\\ MWC 120 & A0 & 7.9 & INT & 1.4& 1.6 & 2460 & 1250 & $\\rm 3.0_{\\rm B},5.0_{\\rm R}$ & $\\rm{1200_B}$,$\\rm{560_R}$&28/12/2008,31/12/2008\\\\ MWC 790 & Be & 12.0 & INT & 3.1 & -- & 4050 &-- & 5.0 &200 & 02/01/2009 \\\\ MWC 137 & Be & 11.2& INT& -- & 2.1 &-- & 4560& 5.0 & 200 & 28/12/2008 \\\\ HD 45677 & $\\rm{Bpshe}$ & 8.0 & WHT & 2.0 & 1.9 &360 &240 &5.0 & $\\rm{900_B}$,$\\rm{550_R}$& 06/10/2006\\\\ LkH$\\rm \\alpha$ 215 & B7.5e & 10.6 & INT & 1.5 & 2.3 &3600 & 3600& $\\rm 5.0_{\\rm B},4.0_{\\rm R}$ & $\\rm{300_B}$,$\\rm{360_R}$ & 27/12/2008,31/12/2008\\\\ MWC 147 & $\\rm{B6pe}$ & 8.8 & WHT & 1.8 & 1.5 & 3000 & 1700 & 5.0 & $\\rm{1200_B}$,$\\rm{500_R}$& 07/10/2006\\\\ MWC 147 & $\\rm{B6pe}$ & 8.8 & INT & 1.3 & -- & 2400 & --& 5.0 & 620& 01/01/2009\\\\ R Mon & B0 & 10.4& INT & 4.2 & 2.5 & 3600 & 3600 & 5.0& $\\rm{100_B}$,$\\rm{240_R}$& 28/12/2008,02/01/2009\\\\ V590 Mon & B8pe & 12.9 & INT & 1.5 & -- & 2670 & --& 5.0 & 200 & 01/01/2009\\\\ V742 Mon & B2Ve & 6.9 & INT & 1.4 & 2.8& 1740 & 2535 & 5.0 & $\\rm{400_B}$,$\\rm{800_R}$ & 30/12/2008,31/12/2008\\\\ OY Gem & Bp[e] & 11.1 & INT &1.8 & -- & 2880 & --& 5. 0 & 100 & 03/01/2009\\\\ GU CMa& $\\rm{B2Vne}$ & 6.6 & WHT & 2.5 & 2.4 & 360 & 360 & 5.0 & $\\rm{1500_B}$,$\\rm{900_R}$&06/10/2006 \\\\ GU CMa & $\\rm{B2Vne}$ & 6.6 & INT &1.8 & -- & 720 &--& 5.0 & 1400 &03/01/2009 \\\\ MWC 166 & B0IVe & 7.0 & WHT & 2.5 & 2.3& 210& 120 &5.0 & $\\rm{1200_B}$,$\\rm{800_R}$& 06/10/2006\\\\ HD 76868 & B5 & 8.0 & INT &1.5 & 2.4 & 4830 & 2100& 5.0 & $\\rm{100_B}$,$\\rm{100_R}$ & 30/12/2008,01/01/2009\\\\ HD 81357 & B8 & 8.4 & INT & 1.7 & -- & 4800 & -- & 5.0& 100 &03/01/2009\\\\ MWC 297 & $\\rm{Be}$ & 12.3 & WHT & 1.4 & 1.2 &4100 & 3120 &5.0 & $\\rm{100_B}$,$\\rm{500_R}$&07/10/2006\\\\ HD 179218 & B9e & 7.2 & WHT & 2.6 & 2.4 & 2100& 1200 &1.0/1.5 & $\\rm{2300_B}$,$\\rm{940_R}$&06/10/2006\\\\ HD 190073 & $\\rm{A2IVpe}$ & 7.8 & WHT & 1.5 & 1.2 &540 & 360&5.0 &$\\rm{600_B}$,$\\rm{800_R}$ &07/10/2006\\\\ BD +40 4124&B2 & 10.7 & WHT & 2.1 & 1.9 & 600 & 660 & 4.0 & $\\rm{500_B}$,$\\rm{370_R}$ &07/10/2006\\\\ MWC 361 & $\\rm{B2Ve}$ & 7.4 & WHT & 1.7 & 1.7 & 1350 & 960 &2.5/4.0 & $\\rm{1400_B}$,$\\rm{1400_R}$ &06/10/2006\\\\ SV Cep & $\\rm{Ae}$& 10.1(\\textit{B}) & INT & 1.6& --& 3000& -- & 5.0 & 600 &02/01/2009\\\\ MWC 655 & B1IVnep & 9.2 & INT & 1.7 & -- & 2400 & -- & 5.0 & 400&03/01/2009\\\\ Il Cep & $\\rm{B2IV/Ve}$ & 9.3 & WHT & 1.4 & 1.2 &3500 & 3000&5.0& $\\rm{800_B}$,$\\rm{500_R}$&07/10/2006\\\\ BHJ 71&B4e & 10.9 & WHT & 1.8 & 1.8 & 1200 & 1080 &4.0 & $\\rm{500_B}$,$\\rm{340_R}$ &06/10/2006\\\\ BHJ 71 & B4e & 10.9 & INT & 1.7 & -- &4200 & --& 5.0 & 500 &01/01/2009\\\\ MWC 1080 & $\\rm{B0}$ & 11.6 & WHT & 2.0 & 2.0 & 3300 & 4170 & 5.0 &$\\rm{200_B}$,$\\rm{500_R}$& 06/10/2006\\\\ \\hline \\end{tabular} \\end{center} \\renewcommand\\footnoterule{} \\footnotetext{{$\\rm^a$ Average seeing in the blue spectral region, approximated by the average of the individual median FWHM, where necessary averaged\\\\ \\hspace*{3.4mm} over multiple slit widths.}} \\footnotetext{{$\\rm^b$ Average seeing in the red spectral region, approximated by the average of the individual median FWHM, where necessary averaged \\\\ \\hspace*{3.4mm}over multiple slit widths.}} \\footnotetext{{$\\rm^c$ \\citet{Theetal1994}}} \\end{minipage} \\end{table*} \\end{center} \\subsection{Data reduction} \\label{data_red} Data reduction was conducted using the Image Reduction and Analysis Facility (IRAF)\\footnote{IRAF is distributed by the National Optical Astronomy Observatories, which are operated by the Association of Universities for Research in Astronomy, Inc., under cooperative agreement with the National Science Foundation \\citep{IRAF}.} and routines written in Interactive Data Language (IDL). Initial data reduction consisted of bias subtraction and flat field division. The total intensity spectra were then extracted from the corrected data in a standard fashion. Wavelength calibration was conducted using the arc spectra, and the wavelength calibration solution had a precision of the order $\\rm{<0.1\\AA}$. \\smallskip Spectroastrometry was performed by fitting Gaussian functions to the spatial profile of the long-slit spectra at each dispersion pixel. This resulted in a positional spectrum, the centroid of the Gaussian as a function of wavelength, and a Full Width at Half Maximum (FWHM) spectrum, the FWHM as a function of wavelength. Spot checks were used to ensure that a Gaussian was an accurate representation of the data. The continuum position exhibited a general trend across the CCD chip: of the order of 10 pixels in the case of the ISIS data and 2 pixels in the IDS data. This was removed by fitting a low order polynomial ($\\rm{4^{th}}$ or $\\rm{5^{th}}$ order) to the continuum regions of the spectrum. \\smallskip All intensity, positional and FWHM spectra at a given PA were combined to make an average spectrum for each PA. A correction for slight changes in the dispersion across PAs was determined by cross-correlating average intensity spectra obtained at different PAs. The correction was then applied to the average intensity, positional and FWHM spectra. The average positional spectra for anti-parallel PAs were then combined to form the average, perpendicular, position spectra, for example: ($\\rm 0^{\\circ} - 180^{\\circ}$)/2 and ($\\rm 90^{\\circ} - 270^{\\circ}$)/2. This procedure eliminates instrumental artifacts as real signatures rotate by $\\rm 180^{\\circ}$ when viewed at the anti-parallel PA, while artifacts remain at a constant orientation. In addition, all positional spectra were visually inspected for artifacts not fully removed by this procedure. As with the positional spectra, the FWHM spectra at anti-parallel PAs were also combined to make two averaged, perpendicular spectra. While FWHM features do not rotate across different PAs, the features observed at anti-parallel PAs were used to exclude artifacts via a visual comparison. All conditions being constant, a real FWHM signature should not change from one PA to the opposite angle at $\\rm +180^{\\circ}$. ", "conclusions": "\\label{conclusions} In this paper we present spectroastrometric observations of a relatively large sample of HAe/Be stars. Here we present the salient findings of this work: \\begin{itemize} \\item{We find a high binary fraction, $\\rm{74\\pm6}~per cent$, consistent with previous studies.} \\item{Using spectroastrometry to separate the unresolved binary spectra we determine spectral types for the components of 9 systems.} \\item{The mass ratios of these systems, determined from the constituent spectral types, are inconsistent with a secondary mass randomly selected from the IMF.} \\item{Although our sample is small this result constrains the mode of binary formation in that the mass ratios and separations of the binary systems observed suggest that the secondary forms via disk fragmentation.} \\item{The properties of the binary systems observed indicate that these systems have not spent a significant amount of time in dense, clustered environments. Therefore, these systems demonstrate that isolated star formation can produce stars as massive as $\\rm{\\sim10-15M_{\\odot}}$.} \\end{itemize}" }, "0910/0910.1632_arXiv.txt": { "abstract": "\\begin{center} \\end{center} We present a type of dark energy models where the particles of dark energy $\\phi$ are dynamically produced via a quantum transition at very low energies. The scale where the transition takes places depends on the strength $g$ of the interaction between $\\phi$ and a relativistic field $\\vp$. We show that a $g\\simeq 10^{-12}$ gives a generation scale $E_{gen} \\simeq 1\\,eV$ with a cross section $\\sigma\\simeq 1 \\,pb$ close to the WIMPs cross section $\\sigma_{w}\\simeq pb$ at decoupling. The number density $n_\\phi$ of the $\\phi$ particles is a source term in the equation of motion of $\\phi$ that generates the scalar potential $v(\\phi)$ responsible for the late time acceleration of our universe. Since the appearance of $\\phi$ may be at very low scales, close to present time, the cosmological coincidence problem can be explained simply due to the size of the coupling constant. In this context it is natural to unify dark energy with inflation in terms of a single scalar field $\\phi$. We use the same potential $v(\\phi)$ for inflation and dark energy. However, after inflation $\\phi$ decays completely and reheats the universe at a scale $E_{RH} \\propto h^2 \\, m_{Pl}$, where $h$ is the coupling between the SM particles and $\\vp$. The field $\\phi$ disappears from the spectrum during most of the time, from reheating until its re-generation at late times, and therefore it does not interfere with the standard decelerating radiation/matter cosmological model allowing for a successful unification scheme. We show that the same interaction term that gives rise to the inflaton decay accounts for the late time re-generation of the $\\phi$ field giving rise to dark energy. We present a simple model where the strength of the $g$ and $h$ couplings are set by the inflation scale $E_I$ with $g=h^2 \\propto E_I/m_{Pl}$ giving a reheating scale $E_{RH} \\propto E_I $ and $\\phi$-generation scale $E_{gen} \\propto E_I^2/m_{pl} \\ll E_{RH}$. With this identification we reduce the number of parameters and the appearance of dark energy is then given in terms of the inflation scale $E_I$. ", "introduction": "The nature and dynamics of Dark Energy ''DE'', which gives the accelerating expansion of the universe at present time, is now days one of the most interesting and stimulating fields of physics. It was discovered more than ten years ago \\cite{first aceleration} and it has been confirmed by further cosmological observations, being now one of the most robust conjecture in modern physics. In fact the acceleration of the present universe has many experimental proofs such as the CMB temperature and fluctuations \\cite{komatsu wmap}, in the matter power spectrum measured by galaxy surveys \\cite{sdss,Tegmark SDSS} and in type Ia supoernovae \\cite{Riess IA supernova,Kowalski supernova,Miknaitis supernova}. The most popular model is the so called $\\Lambda$CDM model, in which a cosmological constant and some amount of cold dark matter are included ''by hand''. Despite its extraordinary consistence with observations, $\\Lambda$CDM is an effective model that leaves many unsolved theoretical question. In fact the existence of the cosmological constant and its order of magnitude have no theoretical justification in $\\Lambda$CDM. The cosmological coincidence problem, that is why the universe is starting to accelerate right now, is also unsolved. Introducing a cosmological constant at the initial stages of the standard cosmological model is specially troublesome since one has to fine tune its value to one parte in $10^{120}$. This problem can be ameliorated if we understand when and how dark energy appears in the universe and this is the main motivation of our present work \\cite{axelfab}. However, our approach will also help us to unify dark energy with inflation. An attractive dark energy alternative to the $\\Lambda$CDM model consists in introducing a ''quintessence'' scalar field $\\phi$ that generates the accelerating expansion \\ci{Q}\\ci{axQ} of the universe due to is dynamics. The dynamics is fixed by its potential $v(\\phi)$ and it is possible to choose potentials that lead to a late time acceleration of the universe \\ci{axQ}. This scalar field can be a fundamental or composite particle as for example bound states \\ci{axDG}. In the second case, the bound quintessence fields are scalar fields composed of fundamental fermions, such as meson fields, and can be generated at low energies as a consequence of a low phase transition scale due to a strong gauge coupling constant \\ci{axDG}. This allows to understand why DE appears at such late times. On the other hand, in the former case the appearance of fundamental scalar field is right at the beginning of the reheated universe and the acceleration of the universe takes place at a much later time due to the classical evolution of the quintessence field $\\phi$. The huge difference in scales between the reheating and dark energy scales requires a fine tuning in the choice of the potential. Here, we present an interesting alternative, namely that the emergence of the fundamental quintessence particles $\\phi$ is originated from a quantum transition taking place at low energies, e.g. as low as $eV$ \\cite{axelfab}. The scale where this transition takes places depends on the strength of the interaction between $\\phi$ and a relativistic field $\\vp$ and it is dynamically determined by the ratio $\\Gamma/H$ where $\\Gamma$ is the transition rate and $H$ the Hubble constant. A value of the coupling $g\\simeq 10^{-12}$ gives a generation scale $E_{gen}\\simeq 1\\,eV$ with a cross section $\\sigma = g^2/32 \\pi E_{gen} \\simeq 1 \\, pb$ close to the WIMPs cross section $\\sigma_{w}\\simeq O(pb)$ at decoupling\\ci{wimp}. The subsequent acceleration of the universe is due to the classical evolution of $\\phi$ due to the scalar potential $v(\\phi)$. Our quantum generation scheme does not aim to derive the potential $v(\\phi)$ but to understand why dark energy dominates at such a late time. Clearly, by closing the gap between the energy today $E_o$, where the subscript $o$ always refers to present time quantities, and that of $\\phi$ production $E_{gen}$, we do not require a fine tuning of the parameters in $v(\\phi)$. Since the appearance of $\\phi$ may be at such low scales this offers a new interpretation and solution to the cosmological coincidence problem in terms of the size of the coupling constant $g$. Furthermore, this late time production of the $\\phi$ particles allows to implement in a natural way a dark energy-inflaton unification scheme. In this scenario, after inflation the field $\\phi$ decays completely and reheats the universe with standard model particles. The universe expands then in a decelerating way dominated first by radiation and later by matter. At low energies the same interaction term that gives rise to the inflaton decay accounts for the quantum re-generation of the $\\phi$ field giving rise to dark energy. In general, it is not complicated to choose a scalar potential such that the universe accelerates in two different regions, at early inflation and dark energy epochs, as in quintessential models \\ci{scalarunification}. However, the universe requires to be most of the time dominated first by standard model relativistic particles and later by matter. The reheating of the universe and the long period of decelerating phase are usually not taken into account in inflation-dark energy unification models and these features are essential in the standard cosmological Big Bang model. In our case, the inflaton-dark energy field is completely absent during most of the time (from reheating until re-generation) and therefore it does not interfere with the standard cosmological model. We will exemplify our inflation-dark energy unified scheme with a simple model. The scalar potential $v(\\phi)$ will have only two parameters fixed by the conditions to give the correct density perturbations $\\delta \\rho/\\rho$ and the present time dark energy scale. The two couplings $h,g$, which give the strength of reheating process with SM particles and the $\\phi$ re-generation process at low energies, respectively, are free parameters but we may take them as $g=h^2 \\propto E_I/m_{pl}$, where $E_I$ is the scale of inflation and it is one of the parameters of $v(\\phi)$. Therefore, starting with four free parameters we can reduce the number to only two and these two are fixed by observations. This gives a reheating scale $E_{RH} \\propto E_I$ and $\\phi$-generation scale $E_{gen}\\propto E_I^2/m_{pl} \\ll E_{RH}$. The paper is organized as follows: in section \\ref{motivations} we give an overview of the late time quantum generation of the quintessence field $\\phi$ and its possible unification with the inflaton field. In section \\ref{regeneration} we present the dark energy quantum generation in detail. In section \\ref{regeneration} we show how to unify inflation and dark energy in terms of a single scalar field in the context of our dark energy quantum generation process and we present a simple model. Finally, in section \\ref{phenomenology} we present the main phenomenological consequences of our model while in section \\ref{conclusions} we resume and conclude. ", "conclusions": "\\label{conclusions} We will now present a summary and conclusions of our work. One of the main goals of this paper was to understand why the dark energy is manifested at such a late time. To achieve this we have presented a novel idea, the quantum generation of dark energy, giving a new interpretation of the late time emergence of DE in terms of a late time quantum production of the quintessence $\\phi$ particles. We take a $2\\leftrightarrow 2$ quantum process between $\\phi$ and a relativistic $\\vp$ particles. The scale where the $\\phi$ field is generated is dynamically determined by the condition $\\Gamma/H=E_{gen}/E\\geq 1$ giving an energy scale $E\\leq E_{gen} $ with $E_{gen} = c_{gen} \\, g^2 \\, m_{pl}$ and $c_{gen} \\simeq 6 \\cdot 10^{-4}$. Therefore the smallness of $E_{gen}$ is due to a small coupling $g$ and for $g\\simeq 10^{-12}$ gives a $E_{gen } \\simeq 1 \\, eV$ and a cross section $\\sigma_{gen}\\simeq 1 pb$. The acceleration of the universe is then due to the classical evolution of $\\phi$ and determined by the scalar potential $v(\\phi)$. We have described in section \\ref{regeneration} a universe that initially contains no $\\phi$ particles, i.e. $n_\\phi=\\Omega_\\phi=\\dot\\phi=v(\\phi)=0$, and once the relativistic particles $\\phi$ are produced they become a source term for the generation of the scalar potential $v(\\phi)$. Once $v(\\phi)$ has been produced the classical equation of motion gives the evolution of $\\phi$. We show in section \\ref{unification} that it is possible to unify inflation and dark energy using the same quintessence field $\\phi$. To achieve the unification we required that the potential $v(\\phi)$ has two flat regions, at high energy for inflation and low energy for dark energy. In this scenario, after inflation the field $\\phi$ decays completely and reheats the universe with standard model particles. The universe expands then in a decelerating way dominated first by radiation and later by matter. At low energies the same interaction term that gives rise to the inflaton decay accounts for the re-generation of the $\\phi$ field giving rise to dark energy. An important difference in the quantum process between $\\phi$ and $\\vp$ at high and low energies is the value of transition rate due to the size of the $\\phi$ mass, $m^2_\\phi(E_I)\\gg m^2_\\phi(E_o)$. We presented in section \\ref{unification} a simple example on how the inflation-dark energy unification can be implemented. We used a potential $v=E_I^4(1-\\arctan[\\phi/f])$ which is flat at high and low energies. The two parameters $E_I, f$ are determined by the density perturbations $\\delta\\rho/\\rho$ and the value of $v_o$ at present time giving $E_I=100\\,TeV, f=10^{-39} eV$. The coupling $g$ between $\\phi$ and $\\vp$ and the coupling $h$ between $\\vp$ and the SM particles are free parameters but can be taken as $g=h^2= q \\, E_I/m_{pl}=(q/100)(E_I/100\\,TeV)10^{-12}$ giving a reheating energy $E_{RH}=(q/100)(E_I/(00\\,TeV)\\, TeV$ and a generation energy $ E_{gen}= (q/100)^2(E_I/100\\,TeV)^2 26 \\, eV$. The cross section between $\\phi$ and $\\vp$ is $\\sigma_{gen} = g^2/32 \\pi E_{gen}^2 \\simeq pb$ quite close to cross section of WIMP dark matter with nucleons $\\sigma_{w} \\simeq pb$ at decoupling. By fixing $g=h^2=q\\,E_I$ we have determined the coupling, which set the scales of reheating and $\\phi$ re-generation, in terms of the inflation scale $E_I$ and we can reduced the number of parameters. Of course this is not the only possible choice of $g$ and $h$.\\\\ To conclude, we have presented a general framework to produce the fundamental quintessence field $\\phi$ dynamically at low energies. The energy scale is fixed by the strength of the coupling and this offers a new interpretation of the cosmological coincidence problem: dark energy domination starts at such small energies because of the size of the coupling constant $g$. Finally, our approach allows for an easy implementation of inflation and dark energy unification with the standard long periods of radiation/matter domination. \\appendix" }, "0910/0910.5756_arXiv.txt": { "abstract": "We present interferometric CO observations of twelve z$\\sim$2 submillimetre-faint, star-forming radio galaxies (SFRGs) which are thought to be ultraluminous infrared galaxies (ULIRGs) possibly dominated by warmer dust (T$_{\\rm dust}\\,\\simgt\\,$40\\,K) than submillimetre galaxies (SMGs) of similar luminosities. Four other CO-observed SFRGs are included from the literature, and all observations are taken at the Plateau de Bure Interferometer (PdBI) in the compact configuration. Ten of the sixteen SFRGs observed in CO (63\\%) are detected at $>$4$\\sigma$ with a mean inferred molecular gas mass of $\\sim$2$\\times$10$^{10}$\\,\\msun. SFRGs trend slightly above the local ULIRG L$_{\\rm FIR}$-L$^\\prime_{\\rm CO}$ relation. Since SFRGs are about two times fainter in radio luminosity but exhibit similar CO luminosities to SMGs, this suggests SFRGs are slightly more efficient star formers than SMGs at the same redshifts. SFRGs also have a narrow mean CO line width, 320$\\pm$80\\,\\kms. Many SMGs have similarly narrow CO line widths, but very broad features ($\\sim$900\\,\\kms) are present in a few SMGs and are absent from SFRGs. SFRGs bridge the gap between properties of very luminous $>$5$\\times$10$^{12}$\\,\\lsun SMGs and those of local ULIRGs and are consistent with intermediate stage major mergers. We suspect that more moderate-luminosity SMGs, not yet surveyed in CO, would show similar molecular gas properties to SFRGs. The AGN fraction of SFRGs is consistent with SMGs and is estimated to be 0.3$\\pm$0.1, suggesting that SFRGs are observed near the peak phase of star formation activity and not in a later, post-SMG enhanced AGN phase. Excitation analysis of one SFRG is consistent with CO excitation observed in SMGs (turning over beyond \\dco). This CO survey of SFRGs serves as a pilot project for the much more extensive survey of {\\it Herschel} and {\\sc SCUBA-2} selected sources which only partially overlap with SMGs. Better constraints on the CO properties of a diverse high-$z$ ULIRG population are needed from ALMA to determine the evolutionary origin of extreme starbursts, and what role ULIRGs serve in catalyzing the formation of massive stellar systems in the early Universe. ", "introduction": "\\label{s:introduction} Ultraluminous infrared galaxies (ULIRGs) exhibit some of the most extreme star formation rates in the Universe. The volume density of higher redshift ULIRGs peaks at z$\\sim$2-3 \\citep{chapman05a}$-$this is also the peak epoch in the cosmic star formation rate density and volume density of active galactic nuclei \\citep[AGN; e.g.][]{fan01a,richards06a}. Not only does this indicate a possible link between supermassive black hole growth and rapid star formation, but it also signals the most active phase in galaxy evolution and formation. ULIRGs exhibit very intense (SFR\\simgt200\\,\\Mpy), short-lived bursts ($\\tau\\,\\sim\\,$100\\,Myr) of star formation. The possible life cycle of a ULIRG, from star-formation dominated, dust-enshrouded galaxy, to obscured AGN and then luminous quasar \\citep[e.g.][]{sanders88b,veilleux09a}, provides a testable evolutionary sequence. The best studied ULIRGs at high redshift are submillimetre galaxies \\citep[SMGs;][]{blain02a} which are characterised by their detection at 850\\um\\, with S$_{850}\\simgt5$\\,mJy. While SMGs put powerful constraints on galaxy evolution theories and the environments of extreme star formation \\citep{greve05a,tacconi06a,tacconi08a,chapman05a,pope06a}, their selection is susceptible to strong temperature biasing \\citep{eales00a,blain04a}. At the mean redshift of radio SMGs, z$\\sim$2.2, observations at 850\\um\\ sample the Rayleigh-Jeans tail of blackbody emission where the observed flux density may be approximated by S$_{850}\\propto$L$_{\\rm FIR}\\,T_{dust}^{-3.5}$. Due to the strong dependence on dust temperature, the 850\\um\\ flux density of warm-dust ULIRGs (T$_{d}\\simgt$40\\,K) might be much lower than cooler dust specimens (T$_{d}=$20-40\\,K) thus causing the warmer-dust galaxies to evade submm detection. This selection bias suggests that a large fraction of the z$\\sim2$ ULIRG population has not been accounted for in current work on high-z star formation. \\citet{chapman04a} describe the first observational effort to identify warm-dust ULIRGs as a population, via the selection of submm-faint radio galaxies (SFRGs) with starburst-consistent rest-UV spectra. While they were thought to be ULIRGs by their similarities to SMGs (similar radio luminosities, optical spectra, stellar masses), without detection in the far-infrared (FIR), there was no direct evidence that their luminosities were in excess of 10$^{12}$\\,\\lsun. Ideally, detection at shorter wavelengths in the infrared, at $\\lambda\\,\\le\\,$500\\,\\um\\, must be used to confirm a ULIRG's luminosity in the absence of submm detection; \\citet{casey09b} used 70\\,\\um\\ detection to confirm that a subset of the SFRG population contains a dominating warmer-dust component with $$\\,=\\,52\\,K. While a population of warm-dust ULIRGs has been shown to exist, some fundamental questions still remain unanswered: are warm-dust ULIRGs in a post-SMG AGN heated phase? Could they be triggered by different mechanisms than the major mergers said to give rise to cold-dust SMGs? Investigating the molecular gas content is fundamental to the characterisation of star formation properties and gas dynamics of a galaxy population. Molecular line transitions from carbon monoxide (CO) are a direct probe of the vast gas reservoirs that are needed to fuel high star formation rates \\citep{frayer99a,greve05a,tacconi06a,tacconi08a,chapman08a}. The gas dynamics which are derived from these observations shed light on galaxies' evolutionary sequences by measuring how disturbed their gas reservoirs are and how long they can maintain their star formation rates with the observed fuel supply. Recent simulations work hint that a ULIRG phase may be triggered by either major merger interactions \\citep[e.g.][]{narayanan09a} or from steady bombardment from low mass fragments \\citep[e.g.][]{dave10a}; linking observations with these different evolutionary scenarios is an essential step in understanding galaxy evolution in the early Universe. In this paper, we present CO molecular gas observations, taken with the IRAM Plateau de Bure Interferometer (PdBI), of twelve SFRGs (and four additional SFRGs from the literature) to compare the population with SMGs and other high redshift star forming galaxies. Section \\ref{s:observations} describes the sample selection, molecular gas observations and ancillary data, while section \\ref{s:results} presents our results, in the form of derived gas and star formation quantities of the SFRG sample. Section \\ref{s:discussion} discusses the gas properties of the sample, compares the population to other high redshift galaxies, and hypothesizes on the role of SFRGs in a broader galaxy evolution context relative to local ULIRGs and SMGs while section \\ref{s:conclusions} concludes. Throughout, we use a $\\Lambda$ CDM cosmology with $H_{\\rm 0} = 71$\\kms~Mpc$^{-1}$, $\\Omega_{\\rm \\Lambda}=0.73$ and $\\Omega_{\\rm m}=0.27$ \\citep{hinshaw09a}. ", "conclusions": "\\label{s:conclusions} We have presented CO molecular gas observations of a sample of submillimetre-faint, star-forming radio galaxies (SFRGs). Due to their non-detection at submillimetre wavelengths and lack of dominant AGN, these ultraluminous, \\uJy\\ radio galaxies are thought to be dominated by star formation but have warmer dust temperatures than SMGs. Out of 16 CO-observed SFRGs (12 from this paper and 4 from the literature), 10 are detected with a mean CO luminosity of $L^\\prime_{\\rm CO[1-0]}\\,\\sim\\,$2.6$\\times$10$^{10}$\\,K\\,km\\,s$^{-1}$\\,pc$^2$, which is slightly less luminous than the CO-observed SMG sample, despite being $\\sim$2$\\times$ less luminous in the radio. We attribute the luminosity difference to a selection bias but suggest that physical driving mechanisms might differ between the very bright ($>$10$^{13}$\\,\\lsun) and moderately bright ($\\sim$10$^{12}$\\,\\lsun) populations. High-resolution radio imaging from MERLIN+VLA shows that the radio emission in the SFRG sample is resolved and extended with mean effective radii $\\sim$2\\,kpc, suggesting that the SFRG radio luminosities are dominated by star formation rather than AGN. The MERLIN+VLA sizes constraints are consistent with similarly analyzed SMG MERLIN+VLA sizes. While we note that AGN do not dominate our sample, it is possible that several of our sources have non-negligible AGN due in part to their selection as radio galaxies. Due to limited FIR data, we use the FIR/radio correlation to derive $L_{FIR}$ and then compute extinction-free star formation rates from the FIR. The star formation efficiencies (SFEs) of SFRGs are comparable within large uncertainties to those of SMGs and local ULIRGs, even though a few sources appear to have very high SFEs \\citep[like those in ][]{chapman08a} or very low SFEs \\citep[like those in ][]{daddi08a}. Those with perceived very high SFEs are more likely AGN dominated than super efficient; their FIR luminosities as calculated from the radio are probably overestimated. SFRGs have narrower CO line widths than the bright subsample of SMGs at the same redshifts ($\\Delta V_{\\rm SFRG}\\,\\sim\\,$320\\,\\kms\\ and $\\Delta V_{\\rm SMG}\\,\\sim\\,$530\\,\\kms), suggesting that SFRGs might have less disturbed dynamical environments. The line width distribution is potentially suggestive of different evolutionary stages or processes between SFRGs and SMGs; however, the observed difference with SMGs could be due to a S/N or luminosity bias. The former would mean that intrinsically broad lines would have underestimated FWHMs due to low S/N. The latter is due to the more thorough spectroscopic sampling of the SMG population. SMGs have higher spectroscopic completeness and also include many objects with AGN signatures in the rest-UV/optical. Any SFRGs which have similar AGN signatures were culled from the sample, thus eliminating some of the potentially brightest SFRGs (most of the bright SMGs which have been surveyed in CO contain optical AGN). While less luminous SMGs (at the same luminosities of SFRGs) exist, few have been observed in CO, thus it is difficult to rule out that CO luminosity might relate to the distribution in CO with line width. Despite selection biases, we have explored the possible physical scenarios triggering warm-dust ULIRGs in contrast to the well studied cold-dust ULIRGs. SFRGs appear to bridge the gap between the properties of $>$10$^{13}$\\,\\lsun SMGs and $\\sim$10$^{12}$\\,\\lsun. Quantitatively, their extended radio emissions suggest sizes consistent with SMGs, implying much larger dynamical masses than local ULIRGs. We show that SFRGs have the same AGN fraction as SMGs and are therefore unlikely to represent a `post-SMG' AGN turn-on phase. Luminous SMGs have been characterised as an early infall stage during a major merger, and local ULIRGs are often described at late-type major mergers. Here, we suggest that SFRGs (and less luminous SMGs) span the range of states during peak merger interaction." }, "0910/0910.4799_arXiv.txt": { "abstract": "{ We estimate sensitivity coefficients to variation of the fine-structure constant $\\alpha$ and electron-to-proton mass ratio $\\mu$ for microwave $\\Lambda$-type transitions in CH molecule and for inversion-rotational transitions in partly deuterated ammonia NH$_2$D. Sensitivity coefficients for these systems are large and strongly depend on the quantum numbers of the transition. This can be used for the search for possible variation of $\\alpha$ and $\\mu$. ", "introduction": "Discrete microwave spectra of molecules are often used for astrophysical studies of possible variation of the fine structure constant $\\alpha=e^2/(\\hbar c)$, the electron-to-proton mass ratio $\\mu=m_\\mathrm{e}/m_\\mathrm{p}$, and the nuclear $g$-factor $g_\\mathrm{n}$. \\cite{Dar03} and \\cite{CK03} pointed out that 18 cm $\\Lambda$-doublet line of OH molecule has high sensitivity to variation of $\\alpha$ and $\\mu$. \\cite{VKB04} have shown that inversion transitions in fully deuterated ammonia $^{15}$ND$_3$ have high sensitivity to $\\mu$-variation, $Q_\\mu=5.6$. According to \\cite{FK07a}, the inversion transition in non-deuterated ammonia has a slightly smaller sensitivity, $Q_\\mu=4.5$. Note that molecular rotational lines have sensitivity $Q_\\mu=1.0$. Because of that, possible variation of constants would lead to apparent velocity offsets between $\\Lambda$-doublet OH line, or ammonia inversion line on one hand and rotational molecular lines, originated from the same gas clouds, on the other hand. This method was used by \\cite{KCL05,FK07a,MFMH08,HMM09} to establish very stringent limits on $\\alpha$- and $\\mu$-variation over cosmological timescale $\\sim 10^{10}$ years. Recently ammonia method was applied by \\cite{LMK08} and \\cite{LML09} to dense prestellar molecular clouds in the Milky Way. These observations provide a bound of a maximum velocity offset between ammonia and other molecules at the level of $|\\Delta V| \\le 28$ m/s. This bound corresponds to $|\\Delta\\mu/\\mu| \\le 3\\times 10^{-8}$, which is two orders of magnitude more sensitive than extragalactic constraints cited above. Taken at face value the measured $\\Delta V$ shows positive shifts between the line centers of NH$_3$ and other molecules and suggests a real offset $\\Delta\\mu/\\mu= (2.2\\pm 0.4_\\mathrm{stat}\\pm 0.3_\\mathrm{sys})\\times 10^{-8}$, see \\cite{LML09}. One of the main possible sources of the systematic errors in such observations is the Doppler noise, i.e. stochastic velocity offsets between different species caused by different spatial distributions of molecules in the gas clouds (see discussions by \\cite{KCL05} and \\cite{LRK08}). Because of that it is preferable to use lines with different sensitivity to variation of fundamental constants of the same species. \\cite{KC04} and \\cite{Koz09} showed that sensitivity coefficients for $\\Lambda$-doublet spectra of OH molecule strongly depend on quantum numbers. In this paper we focus on the CH molecule and on partly deuterated ammonia NH$_2$D. The former is similar to OH and has $\\Lambda$-doublet spectrum, which is highly sensitive to variation of $\\alpha$ and $\\mu$. For the latter the rotational and inversion degrees of freedom are strongly mixed due to the broken symmetry. This leads to a significant variation of the sensitivity of different microwave transitions to $\\mu$-variarion. Note that microwave spectra of CH and NH$_2$D from the interstellar medium were detected by several groups (see, for example, \\cite{RES74,ZT85,OBR85,LRT05,LGR08}, and references therein). ", "conclusions": "We have shown that there are several new microwave lines with high sensitivity to possible variation of the fundamental constants, which have been observed in the interstellar medium. Moreover, one can use the lines of the same species with different sensitivities and significantly reduce the Doppler noise." }, "0910/0910.3493_arXiv.txt": { "abstract": "{The measurement of line broadening in cool stars is in general a difficult task. In order to detect slow rotation or weak magnetic fields, an accuracy of $1$\\,km\\,s$^{-1}$ is needed. In this regime the broadening from convective motion become important. We present an investigation of the velocity fields in early to late M-type star hydrodynamic models, and we simulate their influence on \\element[][]{FeH} molecular line shapes. The M star model parameters range between $\\log{g}$ of $3.0~-~5.0$ and effective temperatures of $2500$\\,K and $4000$\\,K.} {Our aim is to characterize the $\\teff$- and $\\log{g}$-dependence of the velocity fields and express them in terms of micro- and macro-turbulent velocities in the one dimensional sense. We present also a direct comparison between 3D hydrodynamical velocity fields and 1D turbulent velocities. The velocity fields strongly affect the line shapes of \\element[][]{FeH}, and it is our goal to give a rough estimate for the $\\log{g}$ and $\\teff$ parameter range in which 3D spectral synthesis is necessary and where 1D synthesis suffices. Eventually we want to distinguish between the velocity-broadening from convective motion and the rotational- or Zeeman-broadening in M-type stars which we are planning to measure. For the latter \\element[][]{FeH} lines are an important indicator.} {In order to calculate M-star structure models we employ the 3D radiative-hydrodynamics (RHD) code \\texttt{CO$^5$BOLD}. The spectral synthesis on these models is performed with the line synthesis code \\texttt{LINFOR3D}. We describe the 3D velocity fields in terms of a Gaussian standard deviation and project them onto the line of sight to include geometrical and limb-darkening effects. The micro- and macro-turbulent velocities are determined with the ``Curve of Growth'' method and convolution with a Gaussian velocity profile, respectively. To characterize the $\\log{g}$ and $\\teff$ dependence of \\element[][]{FeH} lines, the equivalent width, line width, and line depth are regarded.} {The velocity fields in M-stars strongly depend on $\\log{g}$ and $\\teff$. They become stronger with decreasing $\\log{g}$ and increasing $\\teff$. The projected velocities from the 3D models agree within $\\sim~100$\\,m\\,s$^{-1}$ with the 1D micro- and macro-turbulent velocities. The \\element[][]{FeH} line quantities systematically depend on $\\log{g}$ and $\\teff$.} {The influence of hydrodynamical velocity fields on line shapes of M-type stars can well be reproduced with 1D broadening methods. \\element[][]{FeH} lines turn out to provide a mean to measure $\\log{g}$ and $\\teff$ in M-type stars. Since different \\element[][]{FeH} lines behave all in a similar manner, they provide an ideal measure for rotational and magnetic broadening. } ", "introduction": "Most of our knowledge about stars comes from spectroscopic investigation of atomic or molecular lines. In sun-like and hotter stars, the strength and shape of atomic spectral lines provides information on atmospheric structure, velocity fields, rotation, magnetic fields, etc. Measuring the effects of velocity fields on the shape of spectral lines requires a spectral resolving power between $R \\sim 10,000$ ($\\Delta v = 30$\\,km\\,s$^{-1}$) for rapid stellar rotation, $R \\ga 30,000$ ($\\Delta v = 10$\\,km\\,s$^{-1}$) for slower rotation and high turbulent velocities, and resolution on the order of $R \\sim 100,000$ for the analysis of Zeeman splitting and line shape variations due to slow convective motion. In slowly rotating sun-like stars, usually a large number of relatively isolated spectral lines are available for the investigation of Doppler broadened spectral lines. These lines are embedded in a clearly visible continuum allowing a detailed analysis of individual lines at high precision. At cooler temperature, first the number of atomic lines is increasing so that more and more lines become blended rendering the investigation of individual lines more difficult. At temperatures around 4000\\,K, molecular lines, predominantly \\element[][]{VO} and \\element[][]{TiO}, start to become important. At optical wavelengths, molecular bands in general consist of many lines that are blended so that the absorption mainly appears as an absorption band; individual molecular lines are difficult to identify. At temperatures in the M type stars regime (4000\\,K and less), atomic lines start to vanish because atoms are mainly neutral and higher ionization levels are weakly populated. Only alkali lines appear that are strongly affected by pressure broadening. Thus, the detailed spectroscopic investigation of velocity fields in M dwarfs is very difficult at optical wavelengths. M-type stars emit the bulk of their flux at infrared wavelengths redward of 1\\,$\\mu$m. This implies that observation of high SNR spectra in principle is easier in the infrared. Furthermore, M type stars exhibit a number of molecular absorption bands in the infrared, for example \\element[][]{FeH}. In these bands, the individual lines are relatively well separated and provide a good tracer of stellar velocity fields. The lines are intrinsically much narrower than atomic lines in sun-like stars because Doppler broadening due to the temperature related motion of the atoms and molecules is much reduced. Thus, the lines can be used for the whole arsenal of line profile analysis that has been applied successfully to sun-like stars over the last decades. Examples of analyses using \\element[][]{FeH} lines are the investigation of the rotation activity connection in field M-dwarfs, which requires the measurement of rotational line broadening with an accuracy of $1$\\,km\\,s$^{-1}$ \\citep{2007A&A...467..259R}. Another example is the measurement of magnetic fields comparing Zeeman broadening in magnetically sensitive and insensitive absorption lines \\citep[see e.g.~][]{2006ApJ...644..497R}. A precise analysis of \\element[][]{FeH} lines, however is only possible if the underlying velocity fields of the M dwarfs atmospheres are thoroughly understood. In this paper, we model the surface velocity fields of M type stars and their influence on the narrow spectral lines of \\element[][]{FeH}. We calculate 3D-\\texttt{CO$^5$BOLD} structure models \\citep{2002A&A...395...99L} which serve as an input for the line formation program \\texttt{LINFOR3D} \\citep[based on][]{Bascheck1966}. Turbulence's are included in a natural way using hydrodynamics, so that we are able to investigate the modeled spectral lines for effects from micro- and macro-turbulent velocities in the classical sense and their influence on the line shapes. The comparison with 1D-models gives a rough estimate of the necessity of using 3D-models in the spectral domain of cool stars. In the first part of this paper we investigate the velocity fields in the models and their dependence on $\\log{g}$ and $\\teff$. In the second part, we investigate the influence of velocity fields, $\\log{g}$, and $\\teff$ on the \\element[][]{FeH} molecular lines. ", "conclusions": "We investigated a set of M-star models with $\\teff=2500$\\,K - $4000$\\,K and $\\log{g}=3.0$ -- $5.0$ [cgs]. For these models, the 3D hydrodynamic radiative transfer code \\texttt{CO$^5$BOLD} was used. The horizontal and vertical velocity fields in the 3D models were described with a binning method. The convective turn-over point is clearly visible in the atmospheric velocity dispersion structure. To investigate the influence of these velocity fields on spectral line shapes, a description for geometrical projection and limb-darkening effects was applied. With the use of contribution functions, we took only these parts in the atmosphere into account where the lines were formed. The resulting velocity dispersions range from $400$ -- $1600$\\,m\\,s$^{-1}$ with decreasing $\\log{g}$ and with increasing $\\teff$ from $200$ -- $1400$\\,m\\,s$^{-1}$. These values agree well with velocities deduced from line shapes. We expressed the hydrodynamical velocity fields of the 3D models in terms of the classical micro- and macro-turbulent velocities. With this description and the obtained micro- and macro-turbulent velocities, it is possible to reproduce 3D spectral lines on 1D atmosphere models very accurately, hence time consuming 3D treatment of \\element[][]{FeH} molecular lines in the regime of cool stars is not necessary for line profile analysis. A comparison of our velocities with a set of velocities determined from observations with spectral fitting methods showed that the macro-turbulent velocities agree, but the micro-turbulent velocities are a factor of two or three smaller than the ones determined from observations. A line shift due to the larger up-flowing area in the convection zone was investigated too. It is on the order of a few m/s up to $50$\\,m\\,s$^{-1}$ for a very low gravity model. The time dependent jitter in line positions is only about m/s and would be reduced to \\,mm\\,s$^{-1}$ in a real star, due to high number of contributing elements. In order to use \\element[][]{FeH} molecular lines for investigations of spectroscopic/physical properties in cool stars (e.g. Zeeman- or rotational broadening), we explored the behavior in a set of lines on $\\log{g}$ and $\\teff$. We investigated ten \\element[][]{FeH} lines between 9950\\AA\\ and 9990\\AA\\ on our models with the spectral synthesis code \\texttt{LINFOR3D}. \\element[][]{FeH} lines react on different effective temperatures as expected due to the change in chemical composition and pressure. The lines showed also a weak dependence on surface gravity due to changing densities and pressure. The broadening from velocity fields in the 3D models of the $\\log{g}$ series is very strong, but for the $\\teff$ series the broadening from velocity fields is almost covered by v.d.Waals broadening. The difference in line width for hot models is up to $0.5$\\,km\\,s$^{-1}$ and for low gravity models around $1$\\,km\\,s$^{-1}$. That means for the 1D spectral synthesis, that one has to include correct micro- and macro-turbulent velocities for small surface gravities or hot $\\teff$. Due to the fact that the FWHM $\\log{g}$ dependence of \\element[][]{FeH} lines goes in the opposite direction as the $\\log{g}$ dependence of the velocity fields, the \\element[][]{FeH} lines become a great mean to measure surface gravities in cool stars. Because the velocity fields scale with $\\log{g}$ and it should be easily possible to detect them. \\element[][]{FeH} lines with different quantum numbers do not show significant differences for both, $\\log{g}$- and $\\teff$-series. That means the broadening of the lines does not depend on $J$, $\\Omega$, or the branch. Furthermore lines with weak magnetic sensitivity behave just like lines with strong magnetic sensitivity. All lines are broadened in the same way by thermal and hydrodynamical motions. Only the transition probability expressed in the $\\log{gf}$ value influences the behavior of the lines. The line with the lowest $gf$-values did not saturate at low $\\teff$, but in general they are similar to the other \\element[][]{FeH} lines. It is possible to treat the \\element[][]{FeH} molecular lines with different quantum numbers as homogenous in the absence of magnetic fields. That allows to use \\element[][]{FeH} lines to measure magnetic fields \\citep{2006ApJ...644..497R,2007ApJ...656.1121R}. Hence we conclude, that these lines also are an appropriate mean to measure magnetic field strength in M-type stars." }, "0910/0910.3941_arXiv.txt": { "abstract": "We study for what specific values of the theoretical parameters the axion can form the totality of cold dark matter. We examine the allowed axion parameter region in the light of recent data collected by the WMAP5 mission plus baryon acoustic oscillations and supernovae \\cite{komatsu}, and assume an inflationary scenario and standard cosmology. We also upgrade the treatment of anharmonicities in the axion potential, which we find important in certain cases. If the Peccei-Quinn symmetry is restored after inflation, we recover the usual relation between axion mass and density, so that an axion mass $m_a =(85\\pm3){\\rm~\\mu eV}$ makes the axion 100$\\%$ of the cold dark matter. If the Peccei-Quinn symmetry is broken during inflation, the axion can instead be 100$\\%$ of the cold dark matter for $m_a < 15{\\rm~meV}$ provided a specific value of the initial misalignment angle $\\theta_i$ is chosen in correspondence to a given value of its mass $m_a$. Large values of the Peccei-Quinn symmetry breaking scale correspond to small, perhaps uncomfortably small, values of the initial misalignment angle $\\theta_i$. ", "introduction": "About 84$\\%$ of the non-relativistic matter in the Universe is in the form of cold dark matter (CDM)\\cite{komatsu}, whose nature is still to be discovered. One of the most promising hypothetical candidates that could account for the CDM observed is the axion \\cite{weinberg, peccei}. The properties of the axion as the CDM particle have been studied in various papers \\cite{preskill, beltran}. We examine the possibility that the axion accounts for 100$\\%$ CDM in the light of the WMAP5 mission, baryon acoustic oscillations (BAO) and supernovae (SN) data. We also upgrade the treatment of anharmonicities in the axion potential, which we find important in a specific region of the parameter space. The axion parameter space is described by three quantities, the PQ energy scale $f_a$, the Hubble parameter at the end of inflation $H_I$ and the axion initial misalignment angle $\\theta_i$. ", "conclusions": "" }, "0910/0910.3036_arXiv.txt": { "abstract": " ", "introduction": "\\par\\noindent The cores of Active Galactic Nuclei (AGN), identified as quasars, emit a huge amount of power at visible and ultraviolet frequencies \\cite{bmp}. It obtains its power by the gravitational potential energy of a massive black hole residing at its center \\cite{kori}. The radiation is emitted by the accretion disk surrounding the black hole. In this paper we compute the luminosity of the invisible axions \\cite{PQ1,PQ2,Weinberg,Wilczek,McKay,Kim1,Dine,Zhitnitsky,McKay2,Kim} from AGNs. We also consider a hypothetical light pseudoscalar whose couplings and mass are not related to one another. Our motivation for this study is two folds. The pseudoscalar flux from AGNs may be used to impose limits on its mass and couplings. If the pseudoscalar flux is sufficiently large then it might also provide an explanation for the observed large scale alignment of visible polarizations from quasars. Large scale alignment, on distance scales of a Gpc, has been observed in many regions of the sky \\cite{huts98,huts01,huts02,JNS04}. A statistically significant signal of alignment with the local supercluster has also been observed \\cite{huts98,huts01,huts02,JNS04,JP04}. This effect may be explained in terms of the conversion of photons to pseudoscalars in the local supercluster magnetic field. However this explanation is not consistent with data. The problem arises due to the observed difference in the distribution of polarizations among the Radio Quiet (RQ) and optically selected (O) quasars and the Broad Absorption Line (BAL) quasars. The polarization distribution of the RQ and O quasars peaks at very low values. The magnitude of the mixing required to explain the alignment effect is sufficiently large so as to completely wash out this difference. In Ref. \\cite{JPS02} it was suggested that if the pseudoscalar flux from quasars is sufficiently large at visible frequencies, than conversion in the supercluster magnetic field may consistently explain the alignment with supercluster. In this case the alignment is explained in terms of the conversion of pseudoscalars to photons. We first study the emission of pseudoscalars from the accretion disk via the Compton, Bremsstrahlung and the Primakoff channels. In this calculation we assume the pseudoscalar to be the standard axion. Besides emission from the accretion disk, pseudoscalars may also be produced in the AGN atmospheres due to the conversion of photons to pseudoscalar in the background magnetic field. The probability for this conversion is negligible for the standard axion but can be large if the pseudoscalar mass is very small. ", "conclusions": "\\par\\noindent In this paper, we have found that the luminosity of pseudoscalars from the AGN accretion disk due to Compton, Bremsstrahlung and Primakoff channels is very small in comparison to the photon luminosity. However, the photons in visible and ultraviolet frequencies may convert to pseudoscalars outside the accretion disk due to pseudoscalar-photon mixing in the background magnetic field. Taking extinction of photons into account and using the current limit on the pseudoscalar-photon coupling, we find that the pseudoscalar flux produced by this process is relatively large. For ultraviolet frequencies, which would be observed in the visible range on earth, this flux may dominate the photon flux. A large pseudoscalar flux may provide a consistent explanation for the large scale coherent orientation of the visible polarizations from quasars. \\medskip \\\\" }, "0910/0910.2426_arXiv.txt": { "abstract": "We present results from a multi-month reverberation mapping campaign undertaken primarily at MDM Observatory with supporting observations from around the world. We measure broad line region (BLR) radii and black hole masses for six objects. A velocity-resolved analysis of the H$\\beta$ response shows the presence of diverse kinematic signatures in the BLR. ", "introduction": " ", "conclusions": "" }, "0910/0910.1878_arXiv.txt": { "abstract": "The effect of the chosen analysis energy window on the results of a dark matter experiment is exemplified by the curious intersection of the exclusion plots of the XENON10 and the CDMS experiments. After proving that the narrow energy window XENON10 chose to analyze is indeed the cause of such intersection, a method to determine the high-energy extreme of the recoil energy window an experiment should use is obtained. ", "introduction": " ", "conclusions": "" }, "0910/0910.1094_arXiv.txt": { "abstract": "We use new large area far infrared maps ranging from 65-500\\,$\\mu$m obtained with the AKARI and the Balloon-borne Large Aperture Submillimeter Telescope (BLAST) missions to characterize the dust emission toward the Cassiopeia~A supernova remnant (SNR). Using the AKARI high resolution data we find a new ``\\emph{tepid}'' dust grain population at a temperature of $\\sim35$\\,K and with an estimated mass of 0.06\\,M$_{\\odot}$. This component is confined to the central area of the SNR and may represent newly-formed dust in the unshocked supernova ejecta. While the mass of tepid dust that we measure is insufficient by itself to account for the dust observed at high redshift, it does constitute an additional dust population to contribute to those previously reported. We fit our maps at 65, 90, 140, 250, 350, and 500\\,$\\mu$m to obtain maps of the column density and temperature of ``cold'' dust (near 16~K) distributed throughout the region. The large column density of cold dust associated with clouds seen in molecular emission extends continuously from the surrounding interstellar medium to project on the SNR, where the foreground component of the clouds is also detectable through optical, X-ray, and molecular extinction. At the resolution available here, there is no morphological signature to isolate any cold dust associated only with the SNR from this confusing interstellar emission. Our fit also recovers the previously detected ``hot'' dust in the remnant, with characteristic temperature 100~K. ", "introduction": "\\label{The_CasA_SNR} Determining the origin of cosmic dust is fundamental to our understanding of many astronomical processes, including star formation and galaxy evolution. Galaxies and quasars at high redshift have been found to contain large amounts of dust \\citep[$\\ge10^8$\\,M\\subsun;][]{Dunlop_1994, Archibald_dust_in_distant_quasars_2001, Isaak_2002, Priddey_2008}, at a time when the Universe was only about one tenth of its present age. The main source of dust \\textit{injection} within our Galaxy is thought to be the stellar winds of stars on the asymptotic giant branch (AGB) of the Hertzsprung-Russell (HR) diagram \\citep{morgan_edmunds_2003}. Stars at this early epoch would not have been able to reach the AGB phase in the available time, and therefore cannot be the source of the observed dust. Heavy elements are produced in the explosions of supernovae (SNe) and, for several years, models have predicted that considerable amounts of fresh dust (0.1-1\\,M\\subsun) could also be produced \\citep{Kozasa_1991_SNR_dust_model, Woosley_1995_SNR_dust_model, Clayton_1999_SNR_dust_model, Todini_2001_SNR_dust_model}. The life cycle of high mass stars ($>$8\\,M\\subsun), the progenitors of type II SNe, is sufficiently short for SNe to occur within the required timescales. As a result, SNe have been proposed as a possible solution for the origin of the dust seen at high-redshift. However, in order for SNe to generate sufficient dust mass to fill this gap, each supernova would need to generate 0.4-1\\,M\\subsun~dust \\citep{Dwek_2007}. This quantification does not account for dust grain destruction within the supernova, thereby making it a lower bound. In seeking evidence regarding this hypothesis, the focus has been on dust detectable in SNe and supernova remnants (SNR), as reviewed briefly below. However, it seems less well appreciated that injection is only part of the story. \\citet{Draine_2003} reviews the often-ignored arguments that, at least in our Galaxy, the interstellar dust is continually processed on a timescale $3 \\times 10^8$~yr and most of its mass is formed in the interstellar medium. Nevertheless, injected dust is critical at least as seeds for further evolution of the dust population. Studies of SNe find only trace amounts of dust in the hot ejecta, with typical masses of order 10$^{-4}$\\,M$_{\\odot}$ \\citep{Dwek_1987,Lagage_1996,Arendt_1999,Ercolano_2007,Meikle_2007}. Dust studies in SNR appear more promising, a prime example being Cassiopeia A (Cas~A). Cas~A is the remnant of a type IIb supernova event which occurred around AD 1680 \\citep{Raymond_1984, Thorstensen_2001, Fesen_2006a, Krause_2008_TypeIIb}. The progenitor star is believed to have had a mass greater than 20\\,M\\subsun~\\citep{Perez-Rendon_2002}, and the remnant is at a distance $D \\sim 3.4$\\,kpc \\citep{Reed_1995}. Early observations of Cas~A made with IRAS \\citep{IRAS_1984} and ISO \\citep{ISO_1996} did not extend to longer wavelengths, and therefore detected only the ``hot'' ($\\sim$100\\,K) dust component, whose mass seemed insufficient to provide the levels of dust seen in high redshift galaxies. The low angular resolution also made the study of sub-structure in the remnant at intermediate wavelengths difficult. The \\textit{Spitzer Space Telescope} \\citep{Spitzer_06} has been used to study Cas~A \\citep{Hines2004,Krause_2004_CasA,ennis2006}. Most recently \\cite{Rho_2008}, exploiting the angular and spectral resolution achieved with the \\textit{Spitzer} infrared spectrograph \\citep[IRS;][]{Houck_2004}, find between 0.020 -- 0.054\\,M$_{\\odot}$ of hot dust. This is an order of magnitude greater than that previously measured and they conclude that, within modeling uncertainties for galaxy evolution, this could be sufficient to explain at least the lower limit to the dust levels in high-redshift galaxies presented by \\cite{Isaak_2002}. Much more controversial is the question of a ``cold'' dust component in the SNR, because of the issue of contamination by line-of-sight interstellar emission from dust which is also cold ($\\sim$16\\,K; \\S~\\ref{comp_sep}). \\cite{Dunne_2003_CasA}, using data from the Submillimetre Common User Bolometer Array \\citep[SCUBA;][]{Holland_SCUBA_99}, find evidence for $2-4\\,{\\rm M\\subsun}$ of dust in Cas~A at a temperature of $\\sim$18\\,K, significantly more than the mass of hot dust. Using the same methodology, \\citet{Morgan_2003_Kepler} find $\\sim$1\\,M$_{\\odot}$ of cold dust in Kepler's supernova remnant, a thousand times greater than previous measurements for this SNR. Both of these remnants are sufficiently young for the dust observed to be freshly formed in the remnant, rather than being material swept up from the ISM by the shock-wave \\citep{Dickel_1988, Hughes_1999, Synch_Wright_1999}. More recently, \\cite{Dunne_2009} presented further evidence for cold dust in Cas~A using SCUBA polarization data. These show dust emission polarized in an orientation consistent with that of the magnetic field deduced from the radio synchrotron emission, suggesting the detected dust is in the SNR. They find a conservative lower limit for this dust mass of 1\\,M$_{\\odot}$. They also attribute apparent depolarization at the brightest feature to dilution by line of sight interstellar emission. \\cite{Dwek_2004} has argued that the mass estimates in \\cite{Dunne_2003_CasA} exceed the total anticipated mass ejection of Cas~A and suggest that if the submillimeter observations are valid they might instead imply the presence of a much smaller amount of dust which is a much more efficient millimeter wavelength radiator, such as iron needles. The alignment of such needles could affect the polarization. By correlating cold dust emission with molecular line absorption against the SNR synchrotron emission, \\cite{Krause_2004_CasA} argued that the dust is, in fact, associated with a molecular cloud located along the line of sight to the SNR. Molecular emission has also been studied in this direction (e.g., \\citealt{liszt_lucas_1999}), showing that the cloud(s) extend well beyond the SNR itself. \\citeauthor{Krause_2004_CasA} estimate the fresh dust yield within Cas~A to be at least an order of magnitude lower than that found by \\cite{Dunne_2003_CasA}. Even so, this would still provide the predominant dust mass in Cas~A, bolstering the possibility of explaining the quantities of dust seen at high-redshift. The emission from cold dust peaks in the far-infrared and submillimeter and so neither SCUBA, on the long wavelength side, nor Spitzer, on the short side, is ideal for isolating a cold dust component. Near the thermal peak, the best large scale maps covering Cas~A and its environs are from the Multiband Imaging Photometer for Spitzer \\citep[MIPS;][]{MIPS_2004} at 160\\,$\\mu$m \\citep{Krause_2004_CasA} and from the ISO Serendipity Survey at 170\\,$\\mu$m \\citep{Stickel_2007}. Neither map has full coverage, with MIPS missing data on small scales and ISOSS on larger. We observed Cas~A using the Far-Infrared Surveyor \\citep[FIS;][]{Kawada_2007} instrument on-board AKARI. We obtained fully-sampled images of sub-arcminute resolution covering a wide area surrounding Cas~A in four photometric bands from 50 to 180\\,$\\mu$m. This is the wavelength range over which the emission from newly formed hot dust becomes faint and emission from cold dust begins to dominate. Therefore, these AKARI FIS images are very useful for investigating the presence of colder components of dust in the remnant. The Balloon-borne Large-Aperture Submillimeter Telescope\\footnote{\\tt www.blastexperiment.info} \\citep[BLAST;][]{Pascale_2008} was also used to observe the Cas~A region at 250, 350, and 500\\,$\\mu$m. These bands fill in the cold dust spectral energy distribution (SED) on the long wavelength side of the peak. The high mapping speed of BLAST means it was possible to cover a large area surrounding the SNR, giving data for both the SNR and the interstellar cloud structure in the surrounding region. By contrast, the prior SCUBA maps are of a small area and involve deconvolution of a three-beam chopping pattern which reconstructs large scale power poorly. AKARI and BLAST have the advantage that, like \\textit{Spitzer}, they are not required to perform chopped observations. As a result, our new maps are sensitive to the large scale structure present in the Cas~A field. The relatively large area of these maps, coupled with their sensitivity and wavelength coverage, make them ideal for investigation of cold dust emission from Cas~A and the interstellar clouds. We describe the AKARI and BLAST observations in \\S~\\ref{data}. Several distinct sources of emission are distinguishable in these maps. Taking advantage of the AKARI spatial resolution, in \\S~\\ref{internal_dust} we identify a new morphologically compact source of ``tepid'' ($\\sim$35\\,K) dust emission centered on the SNR whose spectrum is also distinct, peaking between the SED peak of the hot dust in the SNR and that of the cold dust. In \\S~\\ref{photometry} we perform aperture photometry on the maps in the region of the Cas~A SNR. The resulting global SED further illustrates the different spectral components and shows that without the additional morphological information it is not possible to unambiguously distinguish the tepid dust component. In \\S~\\ref{comp_sep} we fit the six-band AKARI-BLAST data with a simple spectral model to make column density and temperature maps for the cold dust. These clearly illustrate the confused nature of cold dust emission on the line of sight to the SNR. In \\S~\\ref{foreground} the derived column density on the line of sight is compared to that obtained by other techniques. We present our conclusions in \\S~\\ref{conclusions}. ", "conclusions": "\\begin{enumerate} \\item{We presented far-infrared/submillimeter data at 65 -- 500\\,$\\mu$m for the Cas~A supernova remnant and the surrounding region. We used these maps to characterize the interstellar dust emission using data from cloud regions well beyond the SNR.} \\item{We used high resolution ARAKI data to probe the spectral region between the hot dust emission from the SNR shock-front and cold dust emission. Using a spectrum-informed clean technique we identified a new tepid dust population at temperature of $\\sim$35\\,K. The mass of this individual dust population was estimated to be 0.06\\,M$\\odot$, but with considerable uncertainty because of its dependence on the choice of $\\kappa$.} \\item{The dust yield for this new and independent tepid component is comparable to that estimated previously for the hot dust component by \\cite{Rho_2008}. While such yields could contribute to the dust masses seen in high redshift galaxies, they are still less than the required level 0.4-1\\,M\\subsun estimated by \\citet{Dwek_2007}. While the mass we measure is insufficient to account for the dust observed at high redshift, when taken in combination with the hot and cold dust masses previously reported by \\cite{Rho_2008} and \\cite{Dunne_2009}, it strengthens the argument for supernovae as a potentially significant source of dust production in the high-redshift universe.} \\item{We developed a simple physically-motivated model of the SED of the SNR and interstellar emission and fit this to six-wavelength bands at each pixel. From this we obtained temperature and dust mass column density maps. The interstellar dust was found to be at a temperature of $\\sim 16.5$\\,K, in keeping with previous measurements, but now better constrained due to the improved wavelength coverage.} \\item{We show that the high level of confusion arising from the interstellar cloud structure projected on the SNR precludes a significant detection of cold dust directly associated with Cas~A. The same source of confusion will have affected previous estimates of cold dust in Cas~A, increasing the uncertainty of those estimates. This analysis was not sufficiently sensitive to identify the lower limiting mass found by \\cite{Dunne_2009} or the lower value by \\cite{Krause_2004_CasA} and therefore does not preclude the possibility of a significant population of cold SNR dust grains with temperature close to that of the interstellar dust. The higher angular resolution data anticipated with \\textit{Herschel} working close to the peak of the cold dust emission, together with correlations with surrogates of the interstellar column density, could result in a more sensitive probe.} \\end{enumerate}" }, "0910/0910.3869_arXiv.txt": { "abstract": "We present infrared interferometric imaging of the S-type Mira star $\\chi$ Cygni. The object was observed at four different epochs in 2005-2006 with the IOTA optical interferometer (H\\ band). Images show up to $40\\%$ variation in the stellar diameter, as well as significant changes in the limb darkening and stellar inhomogeneities. Model fitting gave precise time-dependent values of the stellar diameter, and reveals presence and displacement of a warm molecular layer. The star radius, corrected for limb darkening, has a mean value of $12.1\\,$mas and shows a $5.1\\,$mas amplitude pulsation. Minimum diameter was observed at phase $0.94\\pm0.01$. Maximum temperature was observed several days later at phase $1.02\\pm0.02$. We also show that combining the angular acceleration of the molecular layer with CO ($\\Delta v = 3$) radial velocity measurements yields a $5.9\\pm1.5\\,$ mas parallax. The constant acceleration of the CO molecules -- during 80\\% of the pulsation cycle -- lead us to argument for a free-falling layer. The acceleration is compatible with a gravitational field produced by a $2.1^{+1.5}_{-0.7}$ solar mass star. This last value is in agreement with fundamental mode pulsator models. We foresee increased development of techniques consisting in combining radial velocity with interferometric angular measurements, ultimately allowing total mapping of the speed, density, and position of the diverse species in pulsation driven atmospheres. ", "introduction": "Mira variables are low to intermediate mass AGB stars that pulsate with a period of about 1 year. They have a cool ($T_{\\rm eff}\\leq 3000\\,K$) and extended ($R>100\\,R_\\sun$) photosphere. As such, they are bright ($M_k \\leq -7$) infrared beacons, individually observable far into galaxies of the Local Group \\citep{1999IAUS..191..551Z}. They have the potential to probe places where the distance \\citep[eg, NGC\\,5128 in][]{2004A&A...413..903R} or reddening \\citep[eg, the Galactic Center in][]{2009MNRAS.tmp.1192M} does not allow observation of the fainter/bluer -- and rarer -- Cepheids. However, the challenge to overcome is that Mira stars are both intrinsically complicated and ill-understood. Two important relations are of special interest: the period/luminosity (P/L) and the period/mass/radius (P/M/R). The first relation has been derived from population studies (sequence ``C'' in the LMC from \\citet{2000PASA...17...18W} and also in the globular cluster 47 Tuc from \\citet{2005A&A...441.1117L}). The present best parameterization of the P-L relation within our galaxy is \\citep{2008MNRAS.386..313W}: \\begin{equation} M_k=-(3.51\\pm0.20)(\\log(P)-2.38)-(7.25\\pm0.07)\\,, \\label{eq:PL} \\end{equation} where $P$ is the period in days. The zero point of this relation is the most uncertain parameter, with its dependence on the metallicity hardly known. The main difficulty is that parallax values are inaccurate and error-prone due to the large size and inhomogeneous surface brightness of the objects. The second relation, the P/M/R relation, has more relevance to the fundamental physics of the star. It is extremely dependent on the pulsation mode, but also, less crucially, on the surface density and metallicity. The P/M/R relation has been formally derived from numerical modeling of the fundamental pulsation mode of theses stars \\citep{Wood..89}: \\begin{equation} \\log(P)=-2.07+1.94 \\log(R/R_\\sun)-0.9 \\log(M/M_\\sun)\\,, \\label{eq:PMR} \\end{equation} Twenty years later, this model-derived relationship has still rarely been confronted with observation. This paper is a first step forward to establish the P/L and P/M/R relations on a new firm observational footing. Of the three crucial parameters (distance, mass, and radius), the angular diameter is far from being the easiest value to obtain. Because the surface gravity is several orders of magnitude lower than the sun, the pulsation of Mira variable leads to an extended atmosphere. In the cool upper layers, significant amount of the products of the helium fusion react to form di- and polyatomic molecules including TiO, SiO, CO and H$_2$O. The forest of molecular lines and scattering from the dust lead to exotic intensity distributions not at all like a simple stellar disk. In the past, this substantially affected many stellar angular diameters measurements \\citep{1996AJ....112.2147V,1999A&A...345..221P}, paving the way to contentious discussion on the mode of pulsation \\citep{1998A&A...333..647B,1999ApJ...514L..35Y}. However, nowadays, interferometers are able to provide maps of the brightness distribution as a function of wavelength \\citep{2009A&A...496L...1L,2009ApJ...700..114P}. Images of the Mira star T\\ Lep revealed a shell-like atmosphere, with a bright chromatic zone distinctly detached from the photosphere. This could be the first image of what \\citet{2004A&A...424.1011O} called the MOLsphere, a zone of increased density in which formation of warm molecular species would be favored. Accounting for this layer is the key to obtain a correct value for the diameter \\citep{1999A&A...345..221P,2002ApJ...579..446M,2004A&A...426..279P}. We will also show that we can apply to this layer a modified Baade-Wesselink method to derive the distance and mass of the star. The test star of this paper is $\\chi$ Cyg, a S-type Mira star. It has a pulsation period of 408 days, a photometric magnitude ranging from 5.3 to 13.3, and intense emission lines at postmaximum \\citep{1947ApJ...106..274M}. This suggests a large pulsation amplitude. Images were obtained with the IOTA interferometer at four different stellar phases, chronologically $\\phi = 0.93$, 0.26, 0.69 and 0.79. In the next section, we describe the observations and give a short overview of the dataset. In section~\\ref{sc:imaging}, we use an image reconstruction algorithm to map the brightness distribution of the star. Precise geometrical parameters of the star, including existence of the molecular layer, are determined by model fitting in section~\\ref{sc:param_image}. From these values, temperatures and opacities are deduced in section~\\ref{sec:Physics}. Finally, in section~\\ref{sc:der_ma}, we combine angular acceleration with radial velocities measurements to derive estimations of the distance and mass of the star. ", "conclusions": "" }, "0910/0910.0437_arXiv.txt": { "abstract": "We performed an optical spectroscopic monitoring of the blazar 3C 454.3 from September 2003 to July 2008. Sixteen optical spectra were obtained during different runs, which constitute the first spectroscopic monitoring done in the rest-frame UV region (z=0.859). An overall flux variation of the Mg{\\small II}($\\lambda$2800 \\AA) by a factor $\\sim3$ was observed, while the corresponding UV continuum ($F_{\\rm cont}$ at $\\lambda$3000 \\AA) changed by a factor $\\sim 14$. The Mg{\\small II} emission lines respond proportionally to the continuum variations when the source is in a low-activity state. In contrast, near the optical outbursts detected in 2005 and 2007, the Mg{\\small II} emission lines showed little response to the continuum flux variations. During the monitored period the UV Fe{\\small II} flux changed by a factor $\\sim 6$ and correlated with $F_{\\rm cont}$ ($r=0.92$). A negative correlation between EW(Mg {\\small II}) and $F_{\\rm cont}$ was found, i.e.\\ the so-called ``Intrinsic Baldwin Effect''. ", "introduction": " ", "conclusions": "" }, "0910/0910.2118_arXiv.txt": { "abstract": "The age and chemical composition of the stars in present-day galaxies carry important clues about their star formation processes. The latest generation of population synthesis models have allowed to derive age and stellar metallicity estimates for large samples of low-redshift galaxies. After reviewing the main results about the distribution in ages and metallicities as a function of galaxy mass, I will concentrate on recent analysis that aims at disentangling the dependences of stellar populations properties on environment and on galaxy stellar mass. Finally, new models that predict the response of the full spectrum to variations in [$\\alpha$/Fe] will allow us to derive accurate estimates of element abundance ratios and gain deeper insight into the timescales of star formation cessation. ", "introduction": "The age and chemical composition of the stellar populations in galaxies, together with galaxy mass, are key ingredients to uncover galaxy formation and evolutionary paths. Estimates of stellar populations parameters are derived by interpreting detailed spectral information, such as absorption features, on the basis of stellar population synthesis (SPS) models. Our ability of interpreting galaxy spectra has greatly improved with the development of SPS models that i) predict the full spectrum of simple stellar populations (SSPs) at medium/high resolution, allowing to adjust the models to the data quality rather than the other way round, and ii) have a better coverage of the stellar parameters space, allowing the interpretation of a broader range of stellar populations \\citep{1999ApJ...513..224V,2003MNRAS.344.1000B,2007MNRAS.382..498C}. Together with the development of SPS models and spectral fitting techniques, the statistical power of large spectroscopic surveys has allowed to put on a firm ground our understanding of the stellar populations in nearby galaxies. In this contribution I will briefly review results on the dependence of stellar populations properties on galaxy mass, as obtained from the analysis of Sloan Digital Sky Survey (SDSS) galaxy spectra. I will then discuss to what extent these relations are shaped by the environment in which galaxies reside, as recently analysed in Pasquali et al (2009, submitted) combining the SDSS DR4 catalogue of stellar ages and metallicities with the \\cite{2007ApJ...671..153Y} group catalogue. I will conclude with a brief outlook on new population synthesis models that allow accurate estimates of $\\alpha$-element abundance ratios. ", "conclusions": "" }, "0910/0910.3214_arXiv.txt": { "abstract": "We investigate the quality of associations of astronomical sources from multi-wavelength observations using simulated detections that are realistic in terms of their astrometric accuracy, small-scale clustering properties and selection functions. We present a general method to build such mock catalogs for studying associations, and compare the statistics of cross-identifications based on angular separation and Bayesian probability criteria. In particular, we focus on the highly relevant problem of cross-correlating the ultraviolet GALEX and optical SDSS surveys. Using refined simulations of the relevant catalogs, we find that the probability thresholds yield lower contamination of false associations, and are more efficient than angular separation. Our study presents a set of recommended criteria to construct reliable crossmatch catalogs between SDSS and GALEX with minimal artifacts. ", "introduction": "Astrophysical studies can gain significantly by associating data from different wavelength ranges of the electromagnetic spectrum. Dedicated multi-wavelength surveys have been a strong focus of observational astronomy in recent years, e.g. AEGIS \\citep{Davis_2007}, COSMOS \\citep{Scoville_2007}, or GOODS \\citep{Dickinson_2003}. At redshifts lower than those probed by these surveys, several surveys of NASA's Galaxy Evolution Explorer \\citep[GALEX;][]{Martin_2005} essentially provide the perfect ultraviolet counterparts of the Sloan Digital Sky Survey \\citep[SDSS;][]{York_2000} optical data sets. These surveys or the combination of these datasets enables to provide invaluable insights on stars and galaxy properties. Naturally, these data are taken by different detectors of the separate projects, hence it is required to combine their information by associating the independent detections. Recent work by \\citet{Budavari_2008} laid down the statistical foundation of the cross-identification problem. Their probabilistic approach assigns an objective Bayesian evidence and subsequently a posterior probability to each potential association, and can even consider physical information, such as priors on the spectral energy distribution or redshift, in addition to the positions on celestial sphere. In this paper, we put the Bayesian formalism to work, and aim to assess the benefit of using posterior probabilities over simple angular separation cuts using mock catalogs of GALEX and SDSS. In Section~\\ref{sec_simulations}, we present a general procedure to build mock catalogs that take into account source confusion and selection functions. Section~\\ref{sec_xmatch} provides the details of the cross-identification strategy, and defines the relevant quality measures of the associations based on angular separation and posterior probability. In Section~\\ref{sec_results}, we present the results for the GALEX-SDSS cross-identification, and propose a set of criteria to build reliable combined catalogs. ", "conclusions": "" }, "0910/0910.4328_arXiv.txt": { "abstract": "We made mid-infrared observations of the 10\\,M$_\\odot$ Herbig Be star HD200775 with the Cooled Mid-Infrared Camera and Spectrometer (COMICS) on the 8.2\\,m Subaru Telescope. We discovered diffuse emission of an elliptical shape extended in the north-south direction in $\\sim$1000\\,AU radius around unresolved excess emission. The diffuse emission is perpendicular to the cavity wall formed by the past outflow activity and is parallel to the projected major axis of the central close binary orbit. The centers of the ellipse contours of the diffuse emission are shifted from the stellar position and the amount of the shift increases as the contour brightness level decreases. The diffuse emission is well explained in all of geometry (the shape and the shift), size, and configuration by an inclined flared disk where only its surface emits the mid-infrared photons. Our results give the first well-resolved infrared disk images around a massive star and strongly support that HD200775 is formed through the disk accretion. The disk survives the main accretion phase and shows a structure similar to that around lower-mass stars with 'disk atmosphere'. At the same time, the disk also shows properties characteristic to massive stars such as photoevaporation traced by the 3.4\\,mm free-free emission and unusual silicate emission with a peak at 9.2\\,$\\mu$m, which is shorter than that of many astronomical objects. It provides a good place to compare the disk properties between massive and lower-mass stars. ", "introduction": "In these two decades, many circumstellar disks around young forming/formed stars less than several solar masses are found by direct images in the infrared/visible observations (e.g. McCaughrean \\& O'dell 1996; Fukagawa et al. 2004; Fujiwara et al. 2006; Lagage et al. 2006; see also web database \\footnote{A comprehensive list of spatially resolved disks is available (www.circumstellardisks.org).}). Radio, infrared, and visible line observations confirmed that some disks have rotating disk kinematics as expected (Simon, Dutrey, \\& Guilloteau 2000; Pontoppidan et al. 2008; Acke, van den Ancker, Dullemond 2005). In contrast, the formation scenario for massive stars ($>$8M$_\\sun$) is still unclear. Since, for such massive stars, the time scale for the Kelvin-Helmholtz contraction is shorter than that of free-fall or accretion with accretion rate similar to low mass star cases, they start releasing energy through nuclear fusion even during accretion (Palla \\& Stahler 1993). Then radiation pressure due to their large luminosities may prevent surrounding material from accreting onto the star, in particular, in the case of very massive stars (Kahn 1974; Wolfire \\& Cassinelli 1987). Several ideas are proposed to overcome the problem: mass accretion through circumstellar disks (Yorke \\& Sonnhalter 2002; Krumholz et al. 2009), mass accretion under a much larger accretion rate than that usually considered for low mass stars (McKee \\& Tan 2002; Krumholz et al. 2009), or merging of low mass stars (Bonnell, Vine, \\& Bate 2004). Among these ideas, non-isotropic accretion through their circumstellar disks seems most plausible at present. Such non-isotropic accretion alleviates effective radiation pressure on the accreting material. Supporting evidence for this disk scenario is recent discoveries of rotating gas fragments around possible massive young stellar objects (YSOs) by interferometric observations in the radio, especially in millimeter and submillimeter wavelength regions (Cesaroni et al. 2007). Some of them have a velocity gradient in a direction orthogonal to molecular outflow lobes, which suggests that the gas fragments rotate around the central YSOs (Zhang, Hunter, Sridharan 1998; Cesaroni et al. 2005). About a dozen of objects as candidate disks around massive YSOs of up to 20\\,M$_\\sun$ are found so far (Zhang et al. 1998; Patel et al. 2005; Cesaroni et al. 2006; Beltr\\'{a}n et al. 2006; Cesaroni et al. 2007). Typically, their estimated stellar mass, luminosity, and disk radius are 4 to less than about 20\\,M$_\\odot$, a few $\\times$(10$^3$--10$^4$)\\,L$_\\odot$, and 500--2000\\,AU, respectively. It is suggested that early B Herbig Be stars are surrounded by flattened structures from measurement of depolarization across H$\\alpha$ line although the discussed scale is much smaller (order of several stellar radii) than the disk size indicated above (Vink et al. 2002). While radio interferometric observations have so far been the most successful in unveiling the disk existence around massive YSOs, the resolution around 1$''$ is not sufficient to draw the detailed disk geometry. It is in contrast to the situation that disks around young forming stars less than several solar masses have been well depicted by direct images in visible to infrared wavelengths. For some massive YSOs, existence of disks is discussed from the polarization vector distribution of infrared scattered light image of the outflow cavities (Jiang et al. 2005), but their disks themselves are not seen because such objects are still embedded deeply in their envelopes. Direct images are strongly required to establish the existence and shape of the disks around massive YSOs. Our new approach is to search for disks in the mid-infrared around massive YSOs that have emerged from their natal clouds. Owing to the large luminosities of the central stars, the disk surface can be heated up out to large radii enough to be resolved in the mid-infrared with 8\\,m-class telescopes, which provide 100\\,AU resolution for nearby ($\\sim$400\\,pc) targets. We carry out survey observations for extended emission around Herbig Be stars and report the discovery of a disk around HD200775 in this paper. ", "conclusions": "We have made mid-infrared imaging and spectroscopy of the Herbig B3e star HD200775 of $\\sim$10\\,M$_\\sun$ with Subaru/COMICS. We found elongated emission around the unresolved excess emission in all of the 8.8, 11.7, 18.8, and 24.5$\\mu$m bands. The elongated elliptical emission has a radius of about 1000\\,AU and configuration perpendicular to the CO cavity wall formed by the past outflow and parallel to the projected major axis of the central close binary orbit. The brightness contours of the elliptical emission are shifted from the stellar position or from the unresolved source position. All of these observed characteristics are well explained as the surface emission of the tilted flared circumbinary disk. The present observations provide the first well-resolved infrared images of a disk around 10\\,M$_\\sun$ YSO and strongly support that HD200775 is formed through the disk accretion. The optical depth and the silicate features of the disk emission support the 'disk atmosphere' configuration, which is well modeled for low to intermediate mass stars. We fit the observed image with a flared disk model. The overall observed structure is well explained by the model. The derived flared geometry is much flatter than the disk around the intermediate mass star HD97048. The value of $z(r)/r$ is estimated to be 0.12 to 0.27 for the HD200775 disk, while it is 0.49 for the HD97048 disk. It supports the idea that there is a qualitative difference between disks around massive stars and lower mass stars. The 3.4\\,mm free-free emission image in literature is very similar to the mid-infrared elliptical disk emission in size and shape. It suggests that the disk surface photoevaporates due to the ionizing photons from the central star. The SMA 350\\,GHz observations detect the unresolved emission with a 0.8$''$ beam. The disk mass concentrated around the star is estimated as around 0.02\\,M$_\\sun$, which is similar to that of the minimum mass the solar nebula. The 10\\,$\\mu$m region spectra of the peak unresolved source and at the 1.3$''$ south are discussed. The spectrum of the unresolved source shows 1600\\,K featureless emission, which correspond to the innermost circumstellar disk in the vicinity of the central star(s). The spectrum of the diffuse emission at 1.3$''$ south is dominated by the amorphous silicate emission. The peak at 9.2\\,$\\mu$m is shorter than the usual silicate features and unique to the HD200775 disk. It may suggest alternation of grains due to plasma irradiation." }, "0910/0910.1211_arXiv.txt": { "abstract": "{Recent spectro-polarimetric observations have provided detailed measurements of magnetic field, velocity and intensity during events of magnetic field intensification in the solar photosphere.} {By comparing with synthetic observations derived from MHD simulations, we aim to discern the physical processes underlying the observations, as well as to verify the simulations and the interpretation of the observations.} {We consider the temporal evolution of the relevant physical quantities for three cases of magnetic field intensification in a numerical simulation. In order to compare with observations, we calculate Stokes profiles and take into account the spectral and spatial resolution of the spectropolarimeter (SP) on board Hinode. We determine the evolution of the intensity, magnetic flux density and zero-crossing velocity derived from the synthetic Stokes parameters, using the same methods as applied to the Hinode/SP observations to derive magnetic field and velocity information from the spectro-polarimetric data. } {The three events considered show a similar evolution: advection of magnetic flux to a granular vertex, development of a strong downflow, evacuation of the magnetic feature, increase of the field strength and the appearance of the bright point. The magnetic features formed have diameters of 0.1-0.2\\arcsec. The downflow velocities reach maximum values of 5-10 km/s at $\\tau=1$. In the largest feature, the downflow reaches supersonic speed in the lower photosphere. In the same case, a supersonic upflow develops approximately $200$~s after the formation of the flux concentration. We find that synthetic and real observations are qualitatively consistent and, for one of the cases considered, agree very well also quantitatively. The effect of finite resolution (spatial smearing) is most pronounced in the case of small features, for which the synthetic Hinode/SP observations miss the bright point formation and also the high-velocity downflows during the formation of the smaller magnetic features.} {The observed events are consistent with the process of field intensification by flux advection, radiative cooling, and evacuation by strong downflow found in MHD simulations. The quantitative agreement of synthetic and real observations indicates the validity of both the simulations and the interpretations of the spectro-polarimetric observations.} ", "introduction": "Magnetic field is ubiquitously present in the solar photosphere \\citep{deWijn:etal:2008}. On granular scales, it undergoes continual deformation and displacement. It is swept by the horizontal flows and concentrated in the intergranular lanes. Flows are able to compress the field so that the magnetic energy density $B^{2}/8\\pi$ approaches the kinetic energy density $\\rho v^{2}/2$ of the flow \\citep{Parker:1963,Weiss:1966}. This results in a magnetic field strength of a few hundred Gauss at the solar surface. Further intensification to kG strength is driven by the mechanism referred to as: \\textit{superadiabatic effect} \\citep{Parker:1978}, \\textit{convective collapse} \\citep{Webb:Roberts:1978,Spruit:Zweibel:1979} or \\textit{convective intensification} \\citep{GrossmannDoerth:etal:1998}. The first two concepts are a theoretical idealization of the process. The superadiabatic effect contains the basic idea. \\citet{Parker:1978} pointed out that a thermally isolated dowflowing gas within the flux tube in a superadiabatically stratified environment will be accelerated, which would lead to evacuation of the flux tube. Because of the resulting pressure deficit, the gas inside the flux tube will then be pressed together (together with the frozen-in magnetic field) by the surrounding gas, causing the magnetic pressure to increase until a balance of total pressure (magnetic + gas) is reached. The convective collapse extents the concept to the convective instability. It starts with a flux tube in thermal and mechanical equilibrium with the surrounding hydrostatically superadiabaticly stratified plasma. Since external stratification is convectively unstable, any vertical motion within the flux tube can be amplified. Downward flow will grow in amplitude and drain the material from the flux tube. The process continues until a new equilibrium with a strong field is reached. Different aspects of the concept have been the subject of extensive research \\citep[see][for reviews]{Schuessler:1990,Steiner:1999}. \\begin{figure*} \\centering \\includegraphics[width=0.95\\hsize,]{fig0.ps} \\caption{Maps of the whole simulation domain at $t=140$~s. Normalized continuum intensity at $630$~nm(left), vertical component of magnetic field (middle) and velocity (right) at a geometrical height roughly corresponding to the level of $\\langle\\tau_{500}\\rangle = 1$ ($\\approx 930$~km above the bottom of the computational box).} \\label{fig:snap} \\end{figure*} The term convective intensification is used for magnetic field intensification in realistic MHD simulations, where the process occurs in its full complexity. It is driven by the thermal effect in the surface layer of the magnetic concentration. There, due to the presence of magnetic field, heat transport by convection is reduced. The material inside the concentration radiates more that it receives. This leads to cooling of material which thus starts to sink and partial evacuation of the concentration occurs. Contraction of the magnetic concentration by the surroundings (result of the pressure imbalance) leads to an increase in magnetic field strength. Thus, the simulations \\citep{Nordlund:1983,GrossmannDoerth:etal:1998,Gadun:etal:2001,Voegler:etal:2005,Cheung:etal:2008} bear out the basic properties described by idealized concepts. That is the downflow, the evacuation of the magnetic structure, the field increase and, in some cases, establishment of a new equilibrium. The 3D MHD simulations show that the strong magnetic concentrations form as the horizontal flows in the intergranular lanes advect weak, nearly vertical field and concentrate it at the vertices of granular and mesogranular downflow lanes \\citep{Stein:Nordlund:1998,Stein:Nordlund:2006}. Larger magnetic structures form at sites where a granule submerges and the surrounding field is pushed into the resulting dark region. Whether the formed concentration appears dark or bright in the continuum intensity depends on whether the vertical cooling is compensated or not by the lateral heating due to horizontal energy exchange \\citep{Bercik:etal:2003,Voegler:etal:2005}. This formation scenario is consistent with the observations described by \\citet{Muller:1983} and \\citet{Muller:Roudier:1992}. Their observations show that network bright points form in intergranular spaces, at the junction of converging granules as the magnetic field gets compressed by the converging granular flow. The 2D simulations by \\citet{GrossmannDoerth:etal:1998} revealed that magnetic flux concentrations formed by convective intensification can evolve in different ways. They present two possible outcomes. Depending on the initial magnetic flux, a magnetic concentration can reach a stable state after the process, or can be dispersed due to an upflow that develops as high speed downflowing material rebounces from the dense bottom of the tube. Similar results were presented by \\citet{Takeuchi:1999} and \\citet{Sheminova:Gadun:2000}. Observational evidence was found for both cases. \\citet{BelloGonzalez:etal:2008} reported on the formation of a magnetic feature at the junction of intergranular lanes, without any significant upflow observed. \\citet{BellotRubio:etal:2001}, on the other hand, detected a strongly blueshifted Stokes V profile originating in a upward propagating shock, $13$ minutes after the amplification of magnetic field. \\citet{SocasNavarro:MansoSainz:2005} found that supersonic upflows are actually quite common. Events that are interpreted as convective collapse were detected also with the spectropolarimeter (SP) \\citep{Lites:etal:2001} of the Solar Optical telescope \\citep{Tsuneta:etal:2008} on board Hinode \\citep{Kosugi:etal:2007}. Both, \\citet{Shimizu:etal:2008} and \\citet{Nagata:etal:2008} show cases of high speed downflows followed by magnetic field intensification and bright point appearance. The event described by \\citet{Nagata:etal:2008} shows stronger field strength and upflow in the final phase of evolution. In this paper we give three examples of magnetic field intensification from MURaM simulations and make detailed comparison with the results of \\citet{Nagata:etal:2008} and \\citet{Shimizu:etal:2008}. \\begin{SCfigure*} \\centering \\includegraphics[width=0.75\\textwidth]{fig1_g25l.ps} \\caption{Evolution of the continuum intensity at $630$~nm and the magnetic field in a $4\\arcsec\\times 3\\arcsec$ sized region. \\textit{Left-hand side} (double column): original spatial resolution; red and blue contours outline the downflow of 6~km/s at 80~km and upflow of 4~km/s at 400~km above $\\tau = 1$, respectively; horizontal lines mark the positions of vertical cuts shown in Figs.~\\ref{fig:cut} and ~\\ref{fig:cut_f23}; boxes corresponding to cases I, II and III at coordinates $[2\\arcsec,1.5\\arcsec]$, $[0.5\\arcsec,1\\arcsec]$ and $[3.5\\arcsec,1\\arcsec]$, respectively, enlarged in Fig.~\\ref{fig:evol_zoom}. \\textit{Right-hand side}: synthetic Hinode/SP observations; left column shows continuum intensity while the right column shows the apparent field strength (see the text); red contour marks the region with $0.01$~pm of signal excess (see text); yellow crosses mark the positions of pixels that we studied in detail in Fig.~\\ref{fig:evol_red}. Grey scales cover the range of 0.6-1.6 and 0.8-1.2 for normalized intensity at original and reduced resolution, respectively, $\\pm 1000$~G for the vertical component of magnetic field (original resolution) and $\\pm200$~Mx/cm$^{2}$ for the apparent longitudinal magnetic flux density (Hinode resolution).} \\label{fig:evol} \\end{SCfigure*} ", "conclusions": "Our case study of three examples of magnetic field intensification has shown that in all three cases, the field is advected to the junction of several granules. There, it is confined by converging granular flows, which, in two cases, form a vortex. Owing to the presence of the magnetic field, the thermal effect (radiative cooling) induces evacuation of the flux concentration. The evacuation leads to a downward shift of the optical depth scale within the flux concentrations. The shift is smaller for the smaller features due to the lateral radiative heating, which inhibits further evacuation \\citep{Venkatakrishnan:1986}. As a result, the magnetic field at $\\tau=1$, in the smaller features, is weaker than in the case of the feature with more flux. This is in accordance with recent numerical \\citep{Cheung:etal:2007} and observational \\citep{Rueedi:etal:1992,Solanki:etal:1996} results. During the evacuation, the dowflow velocities reach maximum values of 5-10 km/s at $\\tau=1$. In the case of the largest feature, the downflow extends from the upper boundary of the simulation domain and becomes supersonic in the lower photosphere. The magnetic features formed have diameters of 0.1-0.2~\\arcsec. In the case of the biggest feature, a supersonic upflow develops approximately $200$~s after the formation of the flux concentration. The upflow does not lead to a complete dispersal of the field, but the feature persists until 9 minutes later when it undergoes a fragmentation. The disappearance of the features in all three cases occurs when they meet opposite polarity features, between 3 and 20 minutes after their formation. We also show what happens with the observables when the effects of smearing to observational spatial resolution is taken into account. An important result is that, in the case of small features, Hinode/SP would miss the bright point formation and, in some cases, also the high velocity downflows that develop in the process. On the other hand, the signatures of the evolution of large features are detectable even after the spatial smearing. We show that this case can be quantitatively compared with Hinode/SP observation \\citep{Nagata:etal:2008,Shimizu:etal:2008} and exhibits a very similar evolution. This suggests that the magnetic field intensification process in the MURaM simulations is a faithful description of the process taking place on the Sun. Furthermore, our study indicates that the analysis and interpretation of the observations in terms of the convective intensification process is well-founded." }, "0910/0910.1357_arXiv.txt": { "abstract": "{I present a predictive analysis for the behavior of the far-infrared (FIR)--radio correlation as a function of redshift in light of the deep radio continuum surveys which may become possible using the square kilometer array (SKA). To keep a fixed ratio between the FIR and predominantly non-thermal radio continuum emission of a normal star-forming galaxy, whose cosmic-ray (CR) electrons typically lose most of their energy to synchrotron radiation and Inverse Compton (IC) scattering, requires a nearly constant ratio between galaxy magnetic field and radiation field energy densities. While the additional term of IC losses off of the cosmic microwave background (CMB) is negligible in the local Universe, the rapid increase in the strength of the CMB energy density (i.e. $\\sim(1+z)^{4})$ suggests that evolution in the FIR-radio correlation should occur with infrared (IR;~$8-1000~\\mu$m)/radio ratios increasing with redshift. This signature should be especially apparent once beyond $z\\sim3$ where the magnetic field of a normal star-forming galaxy must be $\\sim$50~$\\mu$G to save the FIR-radio correlation. At present, observations do not show such a trend with redshift; $z\\sim6$ radio-quiet quasars appear to lie on the local FIR-radio correlation while a sample of $z\\sim4.4$ and $z\\sim2.2$ submillimeter galaxies (SMGs) exhibit ratios that are a factor of $\\sim$2.5 {\\it below} the canonical value. I also derive a 5$\\sigma$ point-source sensitivity goal of $\\approx$20~nJy (i.e. $\\sigma_{\\rm RMS} \\sim 4$~nJy) requiring that the SKA specified be $A_{\\rm eff}/T_{\\rm sys}\\approx 15000$~m$^{2}$~K$^{-1}$; achieving this sensitivity should enable the detection of galaxies forming stars at a rate of $\\gtrsim25~M_{\\odot}~{\\rm yr}^{-1}$, such as typical luminous infrared galaxies (i.e. $L_{\\rm IR} \\gtrsim 10^{11}~L_{\\odot}$), at all redshifts if present. By taking advantage of the fact that the non-thermal component of a galaxy's radio continuum emission will be quickly suppressed by IC losses off of the CMB, leaving only the thermal (free-free) component, I argue that deep radio continuum surveys at frequencies $\\gtrsim$10~GHz may prove to be the best probe for characterizing the high-$z$ star formation history of the Universe unbiased by dust. } \\FullConference{Panoramic Radio Astronomy: Wide-field 1-2 GHz research on galaxy evolution\\\\ June 2-5 2009\\\\ Groningen, the Netherlands} \\begin{document} ", "introduction": "Radio continuum emission from galaxies arises due to a combination of thermal and non-thermal processes primarily associated with the birth and death of young massive stars, respectively. The thermal (free-free) radiation of a star-forming galaxy is emitted from H{\\sc ii} regions and is directly proportional to the photoionization rate of young massive stars. Since emission at GHz frequencies is optically thin, the thermal radio continuum emission from galaxies is a very good diagnostic of a galaxy's massive star formation rate. Massive ($\\gtrsim 8~M_{\\odot}$) stars which dominate the Lyman continuum luminosity also end their lives as supernovae (SNe) whose remnants (SNRs) are responsible for the acceleration of cosmic-ray (CR) electrons into a galaxy's general magnetic field resulting in diffuse synchrotron emission. These same massive stars are often the primary sources of dust heating in the interstellar medium (ISM) as their starlight is absorbed and reradiated at far-infrared (FIR) wavelengths by interstellar grains. This common origin between the FIR dust emission and thermal $+$ non-thermal radio continuum emission from galaxies is thought to be the dominant physical processes driving the FIR-radio correlation on global (e.g. \\cite{gxh85,yrc01}, and references therein) and local (e.g. \\cite{ejm06,ejm08}, and reference therein) scales. At present, all indications suggest that the FIR-radio correlation holds out to moderate redshifts (e.g. \\cite{df06,ejm09a,ms09}, and references therein). The detection of large populations of dusty star-forming galaxies at high redshift by ISO and {\\it Spitzer} has underscored the need for reliable star formation rate diagnostics unaffected by dust. While radio emission may provide an excellent advantage over other wavelengths, detecting large populations of high redshift star-forming galaxies at radio wavelengths remains extremely difficult. While detectable with current FIR capabilities, even with a fully operational EVLA, IR-bright star-forming galaxies (e.g. M~82; $L_{\\rm IR} \\approx 4\\times10^{10}~L_{\\odot}$) and moderate LIRGs (i.e. $L_{\\rm IR} \\approx 3\\times10^{11}~L_{\\odot}$) will not be detectable beyond redshifts of $z\\sim 1$ and $z\\sim 2$, respectively. A next-generation radio facility such as the Square Kilometer Array (SKA) should easily remedy this disparity between the depth of FIR and radio continuum surveys. Recently, \\cite{ejmc09} described physically motivated predictions for the evolution of the FIR-radio correlation as a function of redshift arising from variations in the CR electron cooling time-scales as Inverse Compton (IC) scattering off of the Cosmic Microwave Background (CMB) becomes increasingly important. Since the non-thermal component of a galaxy's radio continuum is increasingly suppressed with redshift, radio continuum measurements at moderately high frequency ($\\sim$10~GHz) become one of the cleanest ways to quantify the star formation activity of galaxies at high redshifts unbiased by dust. In this proceedings article I summarize some of these findings with an emphasis placed on what this might mean for deep radio continuum surveys using the SKA. ", "conclusions": "In this proceedings article I have summarized the conclusions of \\cite{ejmc09} which presented a predictive analysis for the expected evolution of the FIR-radio correlation versus redshift arising from variations in the CR electron cooling time-scales as IC scattering off of the CMB becomes increasingly important. In doing so, I have focussed on the value of deep radio continuum surveys in studies of star formation at high-$z$, particularly in the context of the SKA, finding the following: \\begin{enumerate} \\item Deep radio continuum observations at frequencies $\\gtrsim$10~GHz using next generation facilities like the SKA will likely provide the most accurate measurements for the star formation rates of normal galaxies at high~$z$. The non-thermal emission from such galaxies should be completely suppressed due to the increased IC scattering off of the CMB leaving only the thermal (free-free) emission detectable; this situation may be complicated if `anomalous' microwave emission from spinning dust grains \\cite{dl98a} in such systems is not negligible. \\item For normal star-forming galaxies to remain on the local FIR-radio correlation at high redshifts requires extraordinarily large magnetic field strengths to counter IC losses from the CMB. For example, the magnetic field of a $z\\sim6$ galaxy must be $\\gtrsim$500~$\\mu$G to obtain a nominal IR/radio ratio. Thus, galaxies which continue to lie on the FIR-radio correlation at high-$z$, such as the sample of radio-quiet QSO's, most likely have radio output which is dominated by an AGN, or the physics at work is simply unknown. \\item To detect typical LIRGs ($L_{\\rm IR} \\gtrsim 10^{11}~L_{\\odot}$) at all redshifts will require nJy sensitivities at GHz frequencies, specifically a 5$\\sigma$ point-source sensitivity of $\\approx$20~nJy (i.e. $\\sigma_{\\rm RMS} \\approx 4$~nJy). Thus, for the SKA to achieve this sensitivity for a reasonably long (300~hr) integration necessitates that $A_{\\rm eff}/T_{\\rm sys} \\sim 15000$~m$^{2}$~K$^{-1}$. At this sensitivity, the SKA will be sensitive to all galaxies forming stars at $\\sim$25~$M_{\\odot}~{\\rm yr}^{-1}$. This includes a significant amount of sources included in the UV luminosity function work of \\cite{rb07,rb08} between $4\\lesssim z \\lesssim 9$. \\end{enumerate}" }, "0910/0910.0398_arXiv.txt": { "abstract": "{ Supersonic turbulence in molecular clouds is a dominant agent that strongly affects the clouds' evolution and star formation activity. Turbulence may be initiated and maintained by a number of processes, acting at a wide range of physical scales. By examining the dynamical state of molecular clouds, it is possible to assess the primary candidates for how the turbulent energy is injected. } { The aim of this paper is to constrain the scales at which turbulence is driven in the molecular interstellar medium, by comparing simulated molecular spectral line observations of numerical magnetohydrodynamic (MHD) models and molecular spectral line observations of real molecular clouds.} { We use principal component analysis, applied to both models and observational data, to extract a quantitative measure of the driving scale of turbulence. } { We find that only models driven at large scales (comparable to, or exceeding, the size of the cloud) are consistent with observations. This result applies also to clouds with little or no internal star formation activity.} { Astrophysical processes acting on large scales, including supernova-driven turbulence, magnetorotational instability, or spiral shock forcing, are viable candidates for the generation and maintenance of molecular cloud turbulence. Small scale driving by sources internal to molecular clouds, such as outflows, can be important on small scales, but cannot replicate the observed large-scale velocity fluctuations in the molecular interstellar medium.} ", "introduction": "Turbulence is an important agent that controls the evolution (and perhaps formation) of molecular clouds and the subsequent production of stars. As such, it has attracted significant attention from theorists, especially since the advent of numerical supercomputer simulations. Of particular interest is the source(s) of energy injection that create and sustain turbulence in molecular clouds. A number of different mechanisms have been proposed, including supernovae, H{\\sc II} regions, outflows, spiral arms, magneto-rotational instability in galactic disks (Mac Low \\& Klessen 2004; Miesch \\& Bally 1994). These mechanisms may be distinguished by the effective spatial scale at which they preferentially operate, and clues to the nature of the energy injection mechanism(s) may be extracted from spectral line imaging observations of molecular clouds. A number of methods for studying resolved velocity fields in molecular clouds have been developed and applied. These include projected velocity (line centroid) analysis (e.g. Scalo 1984; Miesch \\& Bally 1994; Ossenkopf \\& Mac Low 2002; Brunt \\& Mac Low 2004), the spectral correlation function (SCF; Rosolowsky {\\it et al}{} 1999), velocity channel analysis (VCA; Lazarian \\& Pogosyan 2001, 2004), and principal component analysis (PCA; Heyer \\& Schloerb 1997). To date, these methods have been used to estimate the power law indices of the velocity structure function/power spectrum in molecular clouds from observed data cubes of molecular line emission (e.g. Brunt \\& Heyer 2002b; Heyer \\& Brunt 2004). Application of PCA to Outer Galaxy molecular clouds (Brunt 2003a -- Paper I hereafter) revealed that, in comparison to simple models, the observational record favoured large scale driving of turbulence in the molecular clouds. In their study of the Polaris molecular cloud, Ossenkopf \\& Mac Low (2002) also found that large scale driving of turbulence provided a better explanation of the cloud's velocity structure. In this paper, we construct simulated observations of molecular clouds, derived from computational simulations of interstellar turbulence. The models include magnetic fields and self-gravity and are driven (randomly forced) on a range of spatial scales. We employ PCA to quantitatively investigate the observational signatures of different driving scales. Our numerical measurements are compared to previous PCA results obtained from the simple cloud models of Paper I and to the same measurements made on real molecular clouds. The layout of the paper is as follows. In Section~2, we briefly summarize the PCA method and review the relevant findings of Paper I. Section~3 introduces the numerical models and summarizes the simulated observations of these. In Section~4, we present our results, compare these to corresponding observations, and discuss the implications for the generation of turbulence in molecular clouds. Our conclusions are given in Section~5. ", "conclusions": "}} Both analytical and computational descriptions of turbulence are necessarily constrained by observations of interstellar clouds. A qualitative inspection of Figure~\\ref{fig:pcseq} shows that the eigenimages derived from clouds models with large $\\lambda_{D}$/L$_{c}$ are more consistent with the observations of NGC~7538. More generally, the measured values of $l_{2}$/$l_{1}$ from real molecular clouds are typically $\\gtrsim$~0.2. In Figure~\\ref{fig:lhist} we plot the histogram of $l_{2}$/$l_{1}$ measured in the sample of clouds from Paper I, to which we have added additional measurements from the clouds analyzed in Heyer \\& Brunt (2004). In the combined sample there are 35 clouds in total. Using Figure~\\ref{fig:lolc} as a guide to the relationship between $$ and $\\lambda_{D}$/L$_{c}$, these values imply that the molecular clouds are dominated by turbulence driven on large scales compared to the cloud sizes. This may be simply a result of the driving scale itself determining the size of molecular clouds (Ballesteros-Paredes \\& Mac Low 2002; Paper I). \\begin{figure} \\centering \\includegraphics[width=9cm]{f3.eps} \\caption{A comparison of velocity power spectra (power $P$ versus wavenmuber $k$) obtained from a numerically simulated cloud (HC8, with $k_{d}$~=~7--8) and an fBm field ($k_{cut}$~=~7) of the type used in Paper I to represent turbulent driving at $k$~$\\approx$~7. HC8 has more power at low wavenumbers relative to the fBm field. (The vertical scale in this plot is arbitrary.) } \\label{fig:powerspec} \\end{figure} \\begin{figure} \\centering \\includegraphics[width=9cm]{f4.eps} \\caption{(a) Comparison of $l_{2}$/$l_{1}$ derived from simulated CO observations and observations using the $v$-hist method where opacity and excitation effects are not included. (b) Comparison of $l_{2}$/$l_{1}$ derived from simulated $^{13}$CO and $^{12}$CO observations. For each p lot, the solid line denotes equivalent values along the ordinate and absissca axes. } \\label{fig:isotope} \\end{figure} In our experiment, we have considered the simplified case where a single ``driving scale'' is in operation. Within this limitation we identify large scale driving as the dominant scenario. In reality, turbulence can in principle be driven on multiple scales by a number of mechanisms (Scalo 1987). The origin of large-scale energy injection is discussed by Mac Low \\& Klessen (2004), who concluded that field supernovae were the dominant mechanism in regions where they occur, while magneto-rotational instability (Kim, Ostriker, \\& Stone 2003; Tamburro {\\it et al}{} 2009) may provide a background level. In addition to these, other possible mechanisms include forcing by shocks in spiral arm potentials; Dobbs \\& Bonnell (2007) demonstrate that the scale-dependent velocity dispersion in molecular clouds can be replicated by simulated clouds in a galactic disk with a fixed spiral arm pattern. Most of these processes likely require that the molecular cloud turbulence is inherited from still larger scale motions in the atomic ISM (Elmegreen 1993, Ballesteros-Paredes {\\it et al} 1999; Brunt 2003a). In this scenario, the ``driving'' of molecular cloud turbulence could simply be due to the continuous downward cascade of turbulent energy, that not only injects the turbulence but is also responsible for the (potentially rapid) molecular cloud formation in the first place (Bergin {\\it et al} 2004; Glover \\& Mac Low 2007). The presence of large scale turbulence in molecular clouds would be a natural, inevitable consequence of their formation, and their subsequent evolution can be significantly affected by dynamical events occurring in the larger scale ISM. \\begin{figure} \\centering \\includegraphics[width=9cm]{f5.eps} \\caption{Histogram of $l_{2}$/$l_{1}$ obtained from $^{12}$CO observations of real molecular clouds. The vertical lines mark the mean $l_{2}$/$l_{1}$ derived from the model observations ($^{12}$CO) and the horizontal arrows extend over the range of measured $l_{2}$/$l_{1}$. } \\label{fig:lhist} \\end{figure} Energy injection on (initially) small scales by the spatio-temporally intermittent development of outflows, stellar winds and H{\\sc ii} regions within the cloud may not be well modeled by random forcing methods used in these simulations. These point-like injections of energy can expand their spheres of influence over time and may ultimately contribute to large scale turbulent motions. However, on large scales, these processes are disfavoured on energetic grounds (Mac Low \\& Klessen 2004). While there is evidence that energy injection by outflows can be important over limited scales (e.g. Bally, Devine, \\& Alten 1996; Knee \\& Sandell 2000) it is unlikely that outflow-driven turbulence can explain the origin of molecular cloud turbulence as a whole (Walawender, Bally, \\& Reipurth 2005; Banerjee, Klessen, \\& Fendt 2007). This is demonstrated by recent simulations of outflow-driven turbulence which reveal that energy injection by outflows is not capable of creating turbulence at scales comparable to the cloud size. Models of interacting outflows generated either randomly (Carroll {\\it et al}{} 2008), or self-consistently (Nakamura \\& Li 2007) show that turbulence is only injected with an effective driving scale of about 1/5 to 1/10 the size of the cloud ($\\lambda_{D}/L_{c}$~$\\approx$~0.1--0.2) which is incompatible with our results as summarized in Figure~\\ref{fig:lolc} and Figure~\\ref{fig:lhist}. The observable ratio $l_{2}$/$l_{1}$ is expected to lie in the range 0.02--0.05 when $\\lambda_{D}/L_{c}$~$\\approx$~0.1--0.2, according to our modelling results. Additionally, the cloud modelled by Nakamura \\& Li (2007) is only 1.5~pc in size, and it is unclear whether the effective driving scale would increase (for the same outflow parameterization) if a larger cloud was modelled. If the {\\it fractional} driving scale of 0.1--0.2 is interpreted as a {\\it physical} driving scale of 0.15--0.3~pc, then outflow-driven turbulence would be even less effective in globally exciting turbulence in larger clouds. On the other hand, in larger clouds, more massive and energetic outflows may be expected to be present, but it is not currently clear how (or if) the effective fractional driving scale would increase. An observational estimate of the effective driving scale of turbulence by outflows was found by Swift \\& Welch (2008). They inferred an energy injection scale of 0.05~pc for L1551, which is a small fraction of the the overall cloud diameter of around 1.8~pc. Using this estimate, they found a rough balance between the energy injection rate (from the outflows) and the turbulent dissipation rate, with a characteristic injection/decay timescale of $\\sim$~0.1~Myr, which is substantially less than the inferred cloud age of $\\sim$~4--6~Myr. We note here that some caution is required in interpreting the appearance of injection/decay balance for the outflow-driven turbulence. The dissipation time scale of turbulence is proportional to the driving scale (Mac Low 1999). Swift \\& Welch (2008), in calculating their dissipation rate, used a driving scale of 0.05~pc, and therefore their result shows primarily that the energy injection through outflows is quickly dissipated on short time scales over short length scales. If the L1551 cloud is, or has been, subject to large scale ($\\gtrsim$~1.8~pc) driving of turbulence, then the large scale turbulence is controlled by a much longer dissipation timescale of (1.8/0.05)~$\\times$~0.1~Myr~$\\approx$~3.6~Myr, which is more in line with the cloud age. The small-scale driving of outflows would then occur within the longer time evolution of the cloud, set by the longer dissipation time scales of the initial turbulence, injected on large scales. \\begin{figure} \\centering \\includegraphics[width=9cm]{f6.eps} \\caption{ First and second eigenmages obtained from principal component analysis of the NGC~1333 molecular cloud. The C$^{18}$O data and analysis are confined to the central core region, delineated by the rectangular box on the $^{12}$CO and $^{13}$CO images.} \\label{fig:eigim1333} \\end{figure} The effect of multiple outflows within a small region of space may be seen in the outflow-rich NGC~1333 molecular cloud. Here, Quillen {\\it et al} (2005) describe as many 22 cavities within a 1~pc$^{3}$ volume, possibly excavated by outflow activity, in the $^{13}$CO (J=1--0) map of Ridge {\\it et al} (2003). The cavities have typical diameters of 0.1--0.2~pc, indicative again of a small effective driving scale, as the shells surrounding the cavities would presumably collide and merge at larger scales. However, it is not yet clear that outflows are the driving source for these cavities, as many do not have obvious stellar sources inside them -- see Quillen {\\it et al} (2005) for further discussion. To investigate outflow-driven turbulence from an observational perspective, we applied the PCA method to CO observations of the NGC~1333 molecular cloud. We used the J=1--0 spectral lines of $^{12}$CO and $^{13}$CO observed at FCRAO as part of the COMPLETE project (Ridge {\\it et al} 2006), as well as FCRAO C$^{18}$O J=1--0 spectral line data towards the central core region of NGC~1333. In Figure~\\ref{fig:eigim1333} we show the first two eigenimages obtained from the analysis for each spectral line. For $^{12}$CO and $^{13}$CO, we find ``dipole'' second eigenimage structure characteristic of large scale turbulence, and measure $l_{2}$/$l_{1}$ values of 0.59 ($^{12}$CO) and 0.63 ($^{13}$CO). The overall cloud size is estimated from the $^{12}$CO $l_{1}$ measurement to be 3.27~pc, assuming a distance of 318~pc. These measurements show that turbulence is (or has been) driven on large scales in NGC~1333, and is unlikely to have originated from the outflows, which are confined to the central core region, marked by the small box in Figure~\\ref{fig:eigim1333}. Analysis of the C$^{18}$O data in this box allows us to focus in on the high column density material lying in the immediate vicinity of the outflows. We measure $l_{2}$/$l_{1}$~=~0.18$\\pm$~0.07 for the high column density material traced by C$^{18}$O, which is substantially smaller than the global $l_{2}$/$l_{1}$ values found using $^{12}$CO and $^{13}$CO, but still reasonably consistent with turbulence driven at large scales. Some caution should be applied to this result, because, as noted above, large temporal variations in $l_{2}$/$l_{1}$ can occur in the case of large scale driving. With this proviso, according to our model results, the measured $l_{2}$/$l_{1}$ for the central region would imply a fractional driving scale of $\\lambda_{D}/L_{c}$~$\\approx$~0.5--1.0, or a physical driving scale of 0.43--0.86~pc, based on the measured $l_{1}$~=~0.86~pc for the C$^{18}$O data. For reference, the cavity sizes of 0.1--0.2~pc in NGC~1333, if taken as a measure of the driving scale within the 0.86~pc C$^{18}$O central core region, best match our models driven at $k_{d}$~=~3--4, for which we find $l_{2}$/$l_{1}$~$\\approx$~0.11. Examination of the C$^{18}$O second eigenimage structure reveals that it shares, to some degree, the same north-south ``dipole'' structure seen in the $^{12}$CO and $^{13}$CO second eigenimages. The presence of this signature, along with the $l_{2}$/$l_{1}$~=~0.18 measurement, suggests that both the large-scale turbulence in the cloud as a whole, and small-scale (outflow) driven turbulence are important in this region. The inferred driving scale is therefore likely an intermediate value between that arising from the outflows and that deriving from the large-scale turbulent gradient across the core region. As a caveat, we note that the $^{12}$CO and $^{13}$CO lines are likely to better trace lower density, more spatially extended material than that traced by the C$^{18}$O line, so the relationship between the gradients seen in Figure~\\ref{fig:eigim1333} may not be as obvious as we assume. If the C$^{18}$O gradient is itself caused by outflow activity, then this may indicate an interesting connection between the large and small-scale energy injection mechanisms. \\begin{figure*} \\centering \\includegraphics[width=17cm]{f7.eps} \\caption{ Plots of $\\delta{v}$ versus $l$ using all space and velocity scales obtained from principal component analysis of $^{12}$CO, $^{13}$CO, and C$^{18}$O J=1--0 data in the NGC~1333 moleclar cloud. The solid lines in each plot show the bisector fit to all points; the fitted relationships for $^{12}$CO and $^{13}$CO are repeated as dotted lines in all plots. } \\label{fig:dvl1333} \\end{figure*} To examine the overall scale-dependence of turbulent motions in NGC~1333, in Figure~\\ref{fig:dvl1333} we show plots of $\\delta{v}$ versus $l$ for each isotope from their respective maps (see Brunt \\& Heyer 2000b). It is noteworthy that the $^{12}$CO and $^{13}$CO measurements conform to the typical $\\delta{v}$--$l$ relationship found by Heyer \\& Brunt (2004). Of more interest here is the increase in $\\delta{v}$ seen on scales of $\\sim$~0.1~pc in the C$^{18}$O data, relative to the overall level set by the $^{12}$CO and $^{13}$CO data for the cloud as a whole. This excess kinetic energy likely derives from the effect of outflows in the central core region of the cloud. Although the number of retrieved $\\delta{v}$--$l$ pairs is small, the data are in broad agreement with Quillen et al (2005) who estimate outflow-driven cavity sizes and velocity perturbations of $\\sim$~0.1--0.2~pc and $\\sim$~1~kms$^{-1}$ respectively. Thus internal driving of turbulence can be important in sub-parsec regions of larger clouds, where a large number of outflows can develop. The PCA results for the NGC~1333 cloud as a whole set this in context, revealing the presence of larger-scale turbulence that will evolve on longer time scales than that present in the central core region. Turbulent dissipation in the dense part can therefore be replenished not only by local sources, but by external ``driving'' by larger scale flows originating in the surrounding cloud, as part of the overall hierarchy of turbulent motions. One cannot then consider the central star-forming regions as closed systems, evolving independently of their larger scale surroundings. Our cloud sample as a whole does not support a picture in which large scale turbulent motions have decayed sufficiently so that small scale driving alone is dominant. We conclude that either the clouds are continually driven on large scales, or that most clouds are sufficiently young that the initial seeding of turbulence by the large scale flows that created the cloud has not yet dissipated. Other observations (Ossenkopf \\& Mac Low 2002; Brunt \\& Mac Low 2004) support this conclusion. It is not yet clear whether clouds are continually driven, or whether the turbulence is in a decaying state. Offner, Klein, \\& McKee (2008) find that while their simulated clouds do not readily distinguish between decaying or driven conditions, there is a marginal preference for continual driving. If clouds are driven at large scales, the turbulent dissipation time is comparable to their dynamical time (Mac Low 1999). As noted above, the dipole pattern in the second eigenimage that is observed in all molecular clouds provides an important constraint to candidate driving sources. The dipole reflects the spatial distribution of the largest velocity differences within a cloud. One can not directly discriminate whether these velocity differences are due to shear, compressive, or expanding motions. Large scale driving can readily account for such a pattern as it directly deposits the energy at these scales. Turbulence driven on small scales can in principle provide support on larger scales (Klessen {\\it et al} 2000) and it may be possible for excess small scale energy input to drive large scale expasion motion. More generally, a cluster-forming clump could experience expansion, collapse, or perhaps oscillation about an equilibrium state, depending on how active the star formation is. However, the dipole pattern suggests a more directed flow of material, which would require the combined action of outflows to act in a preferred direction. There is a possible mechanism for outflows to be oriented in a particular direction: a strong magnetic field could result in core collapse along field lines, leading to co-oriented protostellar disks and therefore co-oriented outflows, for which some evidence is presented in Anathpindika \\& Whitworth (2008). The magnetic field strength needed to impose such directivity is likely to inhibit cluster formation, and instead promote star formation in a more distributed, quiescent mode (Heitsch, Mac Low, \\& Klessen 2001; Price \\& Bate 2008). Outflows from newborn stars and H{\\sc ii} regions can also redistribute energy from small to larger scales by driving expanding shells. Such flows may also perturb the magnetic field that threads the molecular cloud to excite Alfv\\'en waves that can further redistribute the outflow energy. However, such activity would again require implausible coherence of location and alignment of outflows to reproduce the observed dipole pattern. Another candidate for driving large scale turbulence ``internally'' is energy injection by H{\\sc ii} regions, as argued by Matzner (2002). However, large scale driving is applicable to molecular clouds where H{\\sc ii} regions are absent, such as G216-2.5 (Maddelena's Cloud; Heyer, Williams, \\& Brunt 2006). So while these mechanisms are no doubt present in some molecular clouds, they cannot explain molecular cloud turbulence in general and their effects will be limited to small scales. If H{\\sc ii} regions become large enough to drive large scale motions, then it is likely that the cloud will be destroyed through photoionization rather than ``driven'' (Matzner 2002; Dale {\\it et al}{} 2005). Turbulence driven at large scales promotes star formation that is clustered, rapid, and efficient, while small scale driving tends to form stars singly, slowly, and inefficiently (Klessen {\\it et al}{} 2000). If the star formation rate can be retarded by (additional) small scale energy injection, it must do this in an environment which can be significantly (perhaps dominantly) influenced by large scale turbulent flows of material. While the large scale versus small scale driving picture can be modified by the effects of magnetic fields (Nakamura \\& Li 2008; Price \\& Bate 2008), it is in a much more dynamic way than that described by the quasistatic model (Shu, Adams, \\& Lizano 1987). For example, recent high spatial dynamic range imaging of the Taurus molecular cloud (Goldsmith {\\it et al}{} 2008; Heyer {\\it et al}{} 2008) reveal large scale, magnetically-regulated, turbulent flows of material. In addition to energy injection, another important consideration is the dissipation of turbulence. Basu \\& Murali (2001) argue that it is difficult to reconcile the inferred heating rate arising from dissipation of turbulence with observed cloud luminosities unless the driving occurs at large scales. More recently, Pan \\& Padoan (2008) show that (assuming large scale driving) heating by turbulent dissipation can exceed cosmic ray heating, and typical temperatures of $\\sim$8.5~K can be sustained by turbulent heating alone. Since the turbulent heating rate scales as $(\\lambda_{D}/L_{c})^{-1}$, widespread small-scale driving could lead to high cloud temperatures that are incompatible with observations for molecular clouds as a whole (although not for small sub-regions within the clouds). We mention a note of caution regarding the results presented here. The numerical simulations of turbulence relied on random forcing (in Fourier space) to generate the turbulent driving, which does not in detail adequately represent many physical sources of energy injection. In the case of outflow-generated turbulence, considered here to be ``small scale'', it was indeed found that the turbulence was effectively driven on small scales. The close correspondence between the numerical models and the simple models of Paper I suggest also that it is not necessarily the details of the flows that are essential, but simply the range of scales on which the turbulence is present. In this sense, the modeling completed so far (Paper I and this work) adequately represent, statistically, turbulence with an outer scale that is detectable in observations. It is to be expected that more realistic driving mechanisms (e.g. as implemented by Nakamura \\& Li 2007) can be investigated in future. Finally, our results recommend that simulations of randomly forced turbulence must necessarily include large scale driving in order to replicate real molecular clouds. How this translates in detail to more realisitic driving mechanisms must be addressed in future work." }, "0910/0910.5268_arXiv.txt": { "abstract": "The development of cosmic ray air showers can be influenced by atmospheric electric fields. Under fair weather conditions these fields are small, but the strong fields inside thunderstorms can have a significant effect on the electromagnetic component of a shower. Understanding this effect is particularly important for radio detection of air showers, since the radio emission is produced by the shower electrons and positrons. We perform Monte Carlo simulations to calculate the effects of different electric field configurations on the shower development. We find that the electric field becomes important for values of the order of 1 kV/cm. Not only can the energy distribution of electrons and positrons change significantly for such field strengths, it is also possible that runaway electron breakdown occurs at high altitudes, which is an important effect in lightning initiation. ", "introduction": "\\label{sec:intro} The effect of atmospheric electric fields on the development of extensive air showers from high energy cosmic rays has not received much attention in the past. Because of the large energies of shower particles, the electric fields present in the atmosphere are generally much too small to alter the particle energies significantly. The largest fields are of the order of 1 kV/cm and only occur in thunderstorms \\cite{book}. In such fields the hadronic and muonic part of the shower are hardly affected, although a muon deficit due to increased decay rate has been reported by Alexeenko et al.~\\cite{A02}. The effects on the electromagnetic shower are much larger, but they are, as we will show in this work, local in the sense that the amount and energy distribution of electromagnetic particles quickly adapts to the local background field. Below thunderstorms the electric field decreases, so particle detector arrays will in general not be strongly sensitive to the influence of electric fields. Mountain top experiments, however, can be very close to, or even inside thunderstorms. An increase in the air shower detection rate and the single particle count has been reported by high altitude experiment such as EAS-TOP \\cite{eastop} and the Carpet air shower array \\cite{A02}. Effects of atmospheric electric fields are generally not included in air shower simulation codes like CORSIKA \\cite{CORSIKA}. There are, however, two topics for which we demonstrate in this paper that it is important to take electric fields into consideration. The first is the consequence of different shower developments in strong electric field regions on the radio emission of the shower. With experiments such as LOPES \\cite{F05} and Codalema \\cite{codalema}, the technique of radio detection of air showers has become mature in the last few years. Radio antennas, with their low costs and high duty cycles, have been demonstrated to be an attractive addition to large air shower arrays like the Pierre Auger Observatory \\cite{PAO}. Already in the 1970s, experimental results by Madolesi et al.~\\cite{M74} showed excessively large radio pulses during thunderstorms. Together with a large range in measured radio intensity and the inability to filter out radio interference, the unknown effect of electrical conditions in the atmosphere was a reason to abandon the radio experiments \\cite{B77} until the development of digital radio arrays like LOFAR \\cite{lofar} revived the interest. A study of the effect of weather conditions on the radio pulse strength with LOPES data confirmed the amplification of radio pulses during thunderstorm conditions, but also showed that no effect is observed under other weather conditions \\cite{B07}. In another study \\cite{N08} it was shown that during thunderstorms the air shower arrival direction reconstructed with the LOPES radio antennas can be a few degrees off with respect to the direction reconstructed by the KASCADE particle detector array. It is now known that the radio emission of air showers can be described in terms of coherent synchrotron radiation of shower electrons and positrons that follow curved trajectories in the Earth's magnetic field as first described in Falcke et al.~\\cite{FG03} and in more detail in Huege et al.~\\cite{HF03}. Alternatively, a macroscopic picture can be constructed in which the shower charges, that are separated in the magnetic field, support a transverse current producing radiation as proposed by Kahn \\& Lerche \\cite{KL66} and more recently by Scholten et al.~\\cite{Sch}. Although there are subtle differences between these descriptions, it is clear that an alteration in the electron and positron distribution of the shower at some altitude, can influence the power of the radio pulse in both models. In this paper we limit ourselves to a study of electric field effects on these distributions, not on the radio pulse itself, which will be the subject of a forthcoming paper. The second issue related to the electric field is the suggestion that air showers of sufficient energy can start an avalanche of runaway electrons in thunderstorm electric fields. Ionization electrons that are produced in collisions of shower particles with air molecules are accelerated in the thunderstorm electric field and can, under the right conditions, gain enough energy to ionize further molecules, an effect described by Gurevich et al.~\\cite{G92}. In thunderstorm research the field strength that can support such avalanches is known as the threshold field, described in Marshall et al.~\\cite{MMR95}. In their work, the authors present thunderstorm measurements which show that lightning often occurs when the thunderstorm field exceeds the breakeven field, suggesting that runaway electron breakdown plays a role in lightning initiation. By providing seed electrons for avalanches, air showers from cosmic rays may play an important role in thunderstorm dynamics. Simulations by Dwyer \\cite{Dw03} have shown that the threshold field strength for an avalanche to develop is slightly higher than the breakeven field, taking into account the effects of elastic scattering. Inside thunderstorms there have been measurements of X-ray bursts \\cite{MP85} and gamma bursts, both from Earth \\cite{G04} and space \\cite{F94}. This emission can be explained in terms of bremsstrahlung emitted by the runaway breakdown electrons \\cite{GM99, Dw08a}. In this work, we simulate the effect of electric fields on the development of air showers with CORSIKA. In Sec.~\\ref{sec:sim} we describe the setup of our simulation and the modification that has been made for CORSIKA. In Sec.~\\ref{sec:effects} we derive some simple analytic estimates and limits to compare with the results. Simulation results are presented in Sec.~\\ref{sec:res} and we conclude with a discussion of the simulation limitations and the consequences for realistic field configurations. ", "conclusions": "We have conducted CORSIKA simulations for vertical and inclined showers of several energies inside electric fields with varying strengths. The evolution of the electromagnetic part of the shower does not change significantly below field strengths of 100 V/cm. For fields of the order of a 1000 V/cm the effect becomes important. Such fields only occur naturally inside thunderclouds. The energy distribution can be altered up to energies of $\\sim 1$~GeV. For positive fields this means positrons can outnumber the electrons over a large energy range, resulting in a positive charge excess. For negative fields electron breakdown is observed for altitudes at which the field is larger than the breakeven field and is most efficient when the field is larger than the threshold field. This electron avalanche is directed antiparallel to the electric field and can geometrically be detached from the shower. The electric field effect on the shower evolution is local in the sense that in a low field region a shower that has traversed a high field region is not much different from a shower that has not. For air showers that traverse thunderstorms this means they are generally affected only in the strong field regions of the cloud. Ground based particle detectors will not be very sensitive to these effects. The radio signal from an air shower that travels through a strong field region can significantly change. It has been established experimentally that under thunderstorm conditions the power of the radio pulse may be much larger than anticipated. The order of magnitude of maximum electric fields that occur in thunderstorms coincide with the field strengths at which our simulations show a considerable change in particle distribution. It should be noted, however, that for the strength of the radio pulse to increase, the shower evolution does not necessarily have to change. The direct acceleration of the particles by the electric field produces radiation just like the magnetic deflection does. The effect of electric fields on the radio emission of air showers will be explored in more detail in a forthcoming paper. Air showers can trigger electron avalanches in regions where the electric field exceeds the threshold field. These avalanches may play a role in lightning initiation. Although electron avalanches can be initiated by any seed electron of sufficient energy in a field exceeding the threshold field, a passing air shower offers the unique scenario in which a huge number of high energy electrons is injected in a very small time. The possible interaction between air showers and thunderstorms can be investigated with a hybrid array of particle detectors and radio receivers. In principle radio antennas can pick up the signal of both air showers and electrical discharges, but it is probably not possible to unambiguously detect an air shower signal behind the violent radiative background of a thunderstorm. With a combination of particle detectors and an array of radio antennas electrical activity after an air shower passage could be imaged, allowing for a study of temporal and spatial coincidences. \\label{sec:con} \\begin{small} % \\vspace*{2ex} \\par \\noindent {\\em Acknowledgements.} Part of this research has been supported by Grant number VH-NG-413 of the Helmholtz Association. \\end{small}" }, "0910/0910.2953_arXiv.txt": { "abstract": "This paper presents CCD observations of the Algol-type eclipsing binaries RZ Dra and EG Cep. The light curves have been analyzed with the PHOEBE software and Wilson-Devinney code (2003 version). A detailed photometric analysis, based on these observations, is presented for both binarity and pulsation. The results indicate semidetached systems where the secondary component fills its Roche lobe. After the subtraction of the theoretical light curve, a frequency analysis was performed in order to check for pulsations of the primary component of each system. Moreover, a period analysis was performed for each case in order to search for additional components around the eclipsing pairs. ", "introduction": "Eclipsing Binaries are important astrophysical objects and several of them are known to contain a pulsating component. It is interesting to study such systems, since extra information can be extracted from both their pulsation and eclipsing properties, leading to a more reliable determination of system parameters.\\\\ EG Cep and RZ Dra are referred as candidate systems for including pulsating components \\citep{S06} and therefore have been selected for observations and study. ", "conclusions": "The LC analysis of EG Cep and RZ Dra showed that both of them are semi-detached systems with the secondary component filling its Roche Lobe. The periodic variations of the orbital periods of these systems could be explained by adopting the existence of a tertiary component, while the steady increase of their period is probably due to the mass transfer procedure. In contrast with the O-C diagram solution, the LC analysis for both systems showed that there is not a third light contribution in the total luminosity of the system. This disagreement can be explained by taking into account the small values of mass of the third body found in each case. Finally, we could not detect any pulsation nature of the primary components of both systems. So, more accurate data (by using larger telescope and better CCD) in the future might reveal possible pulsational behaviour." }, "0910/0910.0167_arXiv.txt": { "abstract": " ", "introduction": " ", "conclusions": "\\begin{figure} \\centering \\vspace{4.25cm} \\special{psfile=bershady_fig13a.c.eps voffset=-5 hoffset=-5 vscale=27 hscale=27} \\special{psfile=bershady_fig13b.c.eps voffset=-5 hoffset=160 vscale=27 hscale=27} \\caption{Total and Specific Grasp versus Spectral Power for a range of instruments on 4m- and 10m-class telescopes (solid and dashed lines, respectively) partially updated from Bershady et al. (2005). See text for comments on instrument efficiency.} \\end{figure} Here we explore the sampled parameter space in spatial versus spectral information, as well as coverage versus resolution, starting with grasp and spectral power (Figure 1.13). Recall that because reliable, consistent measurements of efficiency are unavailable for most instruments, we use grasp instead of etendue ({\\tt warning:} we really want etendue). Note, however, that there is a factor of 6 range in the known efficiencies of instruments tabulated in this Chapter. Further note that there are two ways of viewing the specific grasp. From the perspective of staying photon-limited at high spectral resolution, high specific grasp is important. The ``flip side'' is that low specific grasp implies high angular resolution. Figure 1.14 shows that spatial resolution is higher in NIR instruments, while spectral resolution is higher in optical instruments. Fiber IFUs have the largest specific grasp -- reflected in the bifurcation seen in spatial resolution, i.e., fiber-fed instruments have large footprints per element ($d\\Omega$). There is a trend of decreasing specific grasp going from fiber+lenslet, lenslet, and finally to slicers. ESI has unusually large $A \\ \\times \\ d \\Omega$ for a slicer; RSS in FMS mode has the highest specific grasp overall. \\begin{figure} \\centering \\vspace{5.5cm} \\special{psfile=bershady_fig3.14.c.eps voffset=-12 hoffset=-15 vscale=60 hscale=60 angle=0} \\caption{Spatial resolution (a) and specific grasp (b) versus spectral power for all instruments in Tables 1-4, highlighting differences between optical (filled symbols) and NIR (open symbols), as well as between different coupling methods (labeled).} \\end{figure} Figure 1.14 and 1.15 together show that optical and near-infrared instruments trade spatial resolution for grasp; there are no high-grasp NIR instruments; the highest spectral power instruments are optical. Optical and near-infrared instruments sample comparable total information, with optical instruments sampling a broader range of trades between spatial versus spectral information. Older NIR instruments clearly suffer from being detector-{\\it size} limited. IMACS-IFU stands out as having significantly larger number of total information elements, $N_R \\ \\times \\ N_\\Omega$, and in this sense is on-par with future-generation instruments." }, "0910/0910.2738_arXiv.txt": { "abstract": "{Previous observations with the \\emph{Infrared Astronomical Satellite} and the \\emph{Infrared Space Observatory}, and ongoing observations with \\emph{Spitzer} and \\emph{AKARI} have led to the discovery of over 200 debris disks, based on detected mid- and far infrared excess emission, indicating warm circumstellar dust. In order to constrain the properties of these systems, e.g., to more accurately determine the dust mass, temperature and radial extent, follow-up observations in the submillimetre wavelength region are needed.} {The $\\beta$ Pictoris Moving Group is a nearby stellar association of young (${\\sim}12\\,$Myr) co-moving stars including the classical debris disk star $\\beta$ Pictoris. Due to their proximity and youth they are excellent targets when searching for submillimetre emission from cold, extended, dust components produced by collisions in Kuiper-Belt-like disks. They also allow an age independent study of debris disk properties as a function of other stellar parameters.} {We observed 7 infrared-excess stars in the $\\beta$ Pictoris Moving Group with the LABOCA bolometer array, operating at a central wavelength of 870\\,{$\\mu$}m at the 12-m submillimetre telescope APEX. The main emission at these wavelengths comes from large, cold dust grains, which constitute the main part of the total dust mass, and hence, for an optically thin case, make better estimates on the total dust mass than earlier infrared observations. Fitting the spectral energy distribution with combined optical and infrared photometry gives information on the temperature and radial extent of the disk.} {From our sample, $\\beta$~Pic, HD\\,181327, and HD\\,172555 were detected with at least 3$\\sigma$ certainty, while all others are below 2$\\sigma$ and considered non-detections. The image of $\\beta$~Pic shows an offset flux density peak located near the south-west extension of the disk, similar to the one previously found by SCUBA at the JCMT. We present SED fits for detected sources and give an upper limit on the dust mass for undetected ones.} {We find a mean fractional dust luminosity $\\bar{f}_\\mathrm{dust}=11{\\cdot}10^{-4}$ at $t \\approx 12\\,$Myr, which together with recent data at 100\\,Myr suggests an $f_\\mathrm{dust} \\propto t^{-\\alpha}$ decline of the emitting dust, with ${\\alpha}>0.8$.} ", "introduction": "The first large and unbiased survey of the infrared (IR) sky, conducted by the \\emph{Infrared Astronomical Satellite} (IRAS), revealed that approximately 15$\\%$ of nearby main-sequence stars have an excess of 12--100\\,$\\mu$m emission, corresponding to a luminosity at least $2\\times10^{-5}$ times higher than that expected from a pure stellar photosphere \\citep{Backman1993,Backman1987,Aumann1984}. This indicates that these stars are surrounded by warm circumstellar dust, interpreted as originating in a disk of debris left over from planet formation, which is being heated by stellar optical and ultraviolet radiation, and re-radiating at mid- and far-IR wavelengths. The short orbital lifetime of small dust particles suggests that they are being continuously replenished by collisions of larger bodies \\citep[see e.g.][]{Backman1995}. Deeper and more targeted observations by the \\emph{Infrared Space Observatory} (ISO), \\emph{Spitzer}, and \\emph{AKARI}, have revised and extended this list of debris disk candidates to encompass over 200 stars \\citep{Matthews2007,Rhee2007,DDDB}. Although some of the most nearby debris disks have been imaged in the optical and IR, most are spatially unresolved and characterised solely on fits of the spectral energy distribution (SED) to stellar synthetic spectra and a handful of IR photometry data points. In order to better constrain the SED, and with this the temperature and radial extent of the disk, complementary submillimetre (submm) photometry is needed. Submm observations probe the Rayleigh-Jeans tail of the thermal radiation, where the measured integrated flux is roughly (sometimes with non-negligible correction) proportional to the temperature and the mass of the dust. Since the most efficient submm emitters are large and cool grains, which also dominate the mass of the disk \\citep{Zuckerman2001}, observations with instruments like the Large APEX BOlometer CAmera \\citep[LABOCA,][]{Siringo2007} on the Atacama Pathfinder EXperiment \\citep[APEX,][]{Gusten2006} telescope (operating at 870\\,{$\\mu$}m) or SCUBA \\citep{Holland1999} at the James Clerk Maxwell Telescope \\citep[JCMT,][]{Prestage1996} (operating at 850\\,{$\\mu$}m) provide good estimators for the total dust mass (assuming an optically thin disk at these wavelengths). Another advantage of observations in the submm regime is the ability to investigate cool dust created in very extended debris disks or belts located some hundreds of AU from the star (akin to the Solar system's Kuiper Belt), compared to the inner warm dust at 1--100~AU radius studied in IR surveys. The larger disk region probed by submm, compared to IR, makes it feasible to potentially resolve nearby disks, and morphologically determine outer disk radii, dynamical interaction with unseen planets, etc.~\\citep[e.g.][]{Wyatt2008}, even though the angular resolution in the submm in general is lower than for shorter wavelengths. Young, nearby stars, with a confirmed mid- and far-IR excess are prime targets for a submm search for cold extended disks. The $\\beta$ Pictoris Moving Group \\citep[BPMG,][]{Zuckerman2001a} is a nearby stellar association of 30 identified young ($\\sim12$\\,Myr) co-moving member stars \\citep{Zuckerman2004,Barrado1999}, harbouring the perhaps most studied of all debris disk systems, namely $\\beta$ Pictoris \\citep{Smith1984}, and also e.g., AU~Mic. In this paper we present a study of 7 main-sequence IR excess stars in the BPMG, observed at 870\\,$\\mu$m with LABOCA at APEX in Chile. The stars were selected from the sample of currently known members of BPMG fulfilling the following criteria: (1) located at $\\delta<-50\\degr$, thus, southern stars that are not accessible to JCMT, but that will be observable with ALMA, and (2) previously detected IR excess. However, several of the stars previously observed at 850\\,{$\\mu$}m, e.g.~AT~Mic and AU~Mic, had to be excluded due to time constraints. By targeting members of the same moving group we ensure that any differences in dust properties and mass reflect intrinsic, age-independent, variations, since these stars are assumed to be coeval \\citep{Mentuch2008}. This investigation is the precursor of a larger survey of 10--100~Myr old stellar associations, which will permit a statistical survey of disk properties during crucial time periods of debris disk evolution. The ultimate goal is to determine: the incidence of cold dust disks around main sequence stars; whether these disks can serve as indicators of planetary systems; the physical characteristics, e.g. chemical composition and grain size distribution, as a function of fundamental stellar parameters such as mass, metallicity and age; and the disk lifetimes. ", "conclusions": "We summarize the most important findings of our 870\\,$\\mu$m observations of 7 main-sequence IR-excess stars in the BPMG as follows: \\begin{itemize} \\item[\\textbullet] Out of the 7 stars observed, we made three detections. Two of the detected objects have previously never been detected at submm wavelengths. Our observations increase the frequency of detected submm disks in the BPMG to almost 17\\% (5 out of 30 stars). \\item[\\textbullet] $\\beta$~Pic showed a strong flux density peak centered on the star, with two additional weaker flux peaks, both at a PA of $\\sim36\\degr$ and radial distances of 600--700\\,AU and 2000--2500\\,AU, respectively. The former of these has a position consistent with the ``blob'' imaged by \\citet{Holland1998} at 850\\,$\\mu$m. Both may be background submm galaxies, but surprisingly show a close alignment to the observed optical and IR disk plane. Simple fitting of a stellar blackbody and a disk modified blackbody to the SED data suggests a dust temperature of 89\\,K and a low opacity law exponent $\\beta \\approx 0.7$. We calculate a minimum dust mass using the theoretically derived maximum opacity index for 1\\,mm sized icy grains and arrive at 4.8\\,$M_{\\mathrm{Moon}}$. \\item[\\textbullet] HD\\,181327 is clearly detected but at most marginally resolved with an elongation consistent with the previously imaged inclined circumstellar dust ring. SED fitting gives $T_{\\mathrm{dust}}=70$\\,K and $\\beta=0.15$, but with the 870\\,$\\mu$m flux ending up above the fit, implying the existence of a population of colder dust grains. The calculated minimum dust mass $M_\\mathrm{dust}=34\\,M_{\\mathrm{Moon}}$ is remarkably high. \\item[\\textbullet] HD\\,172555 does not show any significant submm emission at the position of the star, however, in the region between the star and its lower mass companion a very extended flux density distribution is seen. Assuming that the emission comes from dust associated with HD\\,172555 we derive the thermal equilibrium temperature of dust at the projected distances ($\\sim$1000\\,AU), which becomes 15\\,K. The integrated flux in the observed features lies far above what would be expected for a single modified blackbody disk fit to IR photometry data in the SED. Adding a second disk component with a temperature of 10--20\\,K fits nicely with the observed SED, and agrees with the temperature estimated from dust grain distances to the star. The corresponding dust mass would be 10--60\\,$M_{\\mathrm{Moon}}$ and 20--70\\,$M_{\\mathrm{Moon}}$ for the NW and SE feature, respectively. \\item[\\textbullet] We derive the fractional dust luminosity of detected sources, which is then plotted versus stellar spectral type. The mean fractional dust luminosity, $\\bar{f}_\\mathrm{dust}=11{\\cdot}10^{-4}$ measured for debris disk in the BPMG, agrees with the expected stage of collisional dust evolution for such ${\\sim}12\\,$Myr systems predicted from previous observations \\citep[e.g.][]{Su2006,Liu2004,Spangler2001} and models \\citep{Kenyon2008,Wyatt2007}. The large scatter in $f_\\mathrm{dust}$ among co-eval stars could be a sign of stochastic collisional dust production or a consequence of BMPG's evolutionary stage, perhaps just at the onset of collisional dust cascades at $t \\approx 10\\,$Myr. Comparison with data at 100\\,Myr \\citep{Greaves2009} gives $f_\\mathrm{dust} \\propto t^{-\\alpha}$ with ${\\alpha}>0.8$. \\end{itemize}" }, "0910/0910.5118_arXiv.txt": { "abstract": "{ We investigate the number and type of pulsars that will be discovered with the low-frequency radio telescope LOFAR. We consider different search strategies for the Galaxy, for globular clusters and for other galaxies. We show that a 25-day all-sky Galactic survey can find approximately 900 new pulsars, probing the local pulsar population to a deep luminosity limit. For targets of smaller angular size such as globular clusters and galaxies many LOFAR stations can be combined coherently, to make use of the full sensitivity. Searches of nearby northern-sky globular clusters can find new low luminosity millisecond pulsars. Giant pulses from Crab-like extragalactic pulsars can be detected out to over a Mpc. } ", "introduction": "Since the discovery of the first four pulsars with the Cambridge radio telescope \\citep{hbp+68}, an ongoing evolution of telescope systems has doubled the number of known radio pulsars roughly every 4 years: An evolution from a large flat receiver with a fixed beam on the sky (the original Cambridge radio telescope) to focusing dishes \\citep[Arecibo --][]{ht75b}, often steerable (Green Bank Telescope), on both hemispheres \\citep[the Parkes telescope --][]{mlc+01}, with large bandwidths and multiple simultaneously usable receivers for wider fields of view (Parkes, Arecibo). The types of pulsars discovered have changed accordingly, from slow, bright, single and nearby pulsars (the original four) to fast (young and millisecond pulsars), far-away (globular clusters) or dim pulsars, some of which are in binaries. The next step in radio telescope evolution will be the use of large numbers of low-cost receivers that are combined interferometrically. These telescopes, the Allen Telescope Array \\citep{bow07b}, LOFAR \\citep{rot03}, MeerKAT \\citep{jon07}, ASKAP \\citep{jtb+08} and the SKA \\citep{kram04}, create new possibilities for pulsar research. In this paper, we investigate the prospects of finding radio pulsars with LOFAR, the LOw Frequency ARray. We outline and compare strategies for targeting normal and millisecond pulsars (MSPs), both in the disk and globular clusters of our Galaxy, and in other galaxies. ", "conclusions": "Because of its large area and field of view, LOFAR can reveal the local population of pulsars to a very deep luminosity limit. A 25-day all-sky survey at 140\\,MHz would find 900 new pulsars, disclosing the local low-luminosity population and roughly doubling the number of pulsars known in the northern hemisphere. Millisecond pulsars in nearby globular clusters can be detected to lower flux limits than previously possible. Assuming the pulsar population in other galaxies is similar to that in ours, we can detect periodicities or giant pulses from extragalactic pulsars up to several Mpcs away." }, "0910/0910.2134_arXiv.txt": { "abstract": "% We compare the luminosity function and rate inferred from the GBM long bursts peak flux distribution with those inferred from the Swift and BATSE peak flux distribution. We find that the GBM, BATSE and the Swift peak fluxes can be fitted by the same luminosity function implying the consistency of these three samples. Using the trigger algorithm of the LAT instrument we derive important information on the flux at 100 MeV compared to lower energy detected by the GBM. We find that the simple extension of the synchrotron emission to high energy cannot justify the low rate of GRBs detected by LAT and for several GRBs detected by the GBM, the flux at $>100$ MeV should be suppressed. Two bursts, GRB090217 and GRB 090202b, detected by LAT have very soft spectra in the GBM and therefore their high energy emission cannot be due to an extension of the synchrotron. ", "introduction": "Gamma ray bursts (GRBs) are the most powerful events in the universe, their total emitted energy outputs exceeding sometime $10^{54}$ ergs, owing to relativistic aberration. In GRBs in fact, the most extreme relativistic regimes are attained among all high energy sources of macroscopic size. Most of the energy is radiated in gamma-rays of 100 to 1000 keV, with tails up to the GeV domain, as detected formerly by CGRO/EGRET (Dingus 1995, Hurley et al. 1994, Gonzalez et al. 1994) and more recently and with much better detail by the AGILE/GRID and Fermi/LAT instruments (Marisaldi et al. 2009, Giuliani et al. 2008, Giuliani et al. 2009, Abdo et al. 2009, Abdo et al. 2009, Bissaldi et al. 2009, Granot 2009). Long GRBs ($T_{90}>2$s) are thought to trace the history of massive star formation of the Universe and are detected all the way from locally (40 Mpc) to the edge of the Universe ($z \\sim 8$, Salvaterra et al. 2009; Tanvir et al. 2009), so that they are ideal targets for the unbiased study of cosmological effects on the propagation of high energy photons and on the evolution of star formation. However the number of GRBs with a measured redshift is still limited and at present we cannot derive directly the GRB luminosity function and rate evolution that are fundamental to understand the nature of these objects. We can constrain the luminosity function and rate distribution by fitting their peak flux distributions to those expected for a given luminosity function and GRB rate, as done for CGRO/BATSE and {\\it Swift}/BAT GRBs (Piran 1992, Cohen \\& Piran 1995, Fenimore \\& Bloom 1995, Loredo \\& Wasserman 1995, Horack \\& Hakkila 1997, Loredo \\& Wasserman 1998, Schmidt 1999, Schmidt 2001, Sethi \\& Bhargavi 2001, Guetta, Piran \\& Waxman 2005, Guetta \\& Piran 2005, 2006, 2007). Since the observed flux distribution is a convolution of these two unknown functions we must assume one and find a best fit for the other. Here we concentrate on the $\\sim$ 250 long GRBs detected so far by the Gamma Ray Burst Monitor (GBM) onboard the Fermi GST (Table 1). We assume that the rate of long bursts follows the star formation rate and we search for the parameters of the luminosity function. In the first part of this paper, we show that one can obtain a fully consistent fit for the GBM, BATSE and the {\\it Swift} peak flux populations, implying that the GBM sample has properties similar to those of BATSE and {\\it Swift}. Only 10 long bursts have been detected above 30 MeV by the LAT. Considering only the ratio between the LAT and GBM field of view (2.5 sr and more than 8 sr, respectively) and assuming comparable sensitivity of the 2 instruments in their energy ranges to the level of emission expected from GRBs, we would have expected that $\\sim 1/3$ of the GBM bursts has a $> 30$ MeV counterpart and not only $\\sim$ 5\\%, as detected. There could be 2 effects at play: the LAT sensitivity and the fact that the intrinsic brightness of the very high energy tails of GRBs can be lower than predicted with a simple extrapolation of the soft gamma-ray spectrum or based on synchrotron and inverse Compton flux estimates (see e.g. Ando, Nakar \\& Sari 2008, who have predicted a LAT detection rate of about 20 GRB/yr and see Fan 2009). Another possible explanation not related to instrumental or intrinsic GRB physics reasons is that in very distant GRBs with highly collimated jets, the intrinsic strong high energy emission is suppressed by the diffuse extragalactic infrared background (Gilmore, Prada, \\& Primack 2009). The statistics of the LAT detection with respect to GBM detection is not dissimilar from that of the GRO instruments EGRET vs BATSE: if the detection were only related to the FOV extension, EGRET should have detected about 50 GRBs out of the nearly 3000 detected by BATSE (BATSE covered virtually all sky, while the EGRET FOV was 30 deg in diameter). The fact that EGRET detected only about 20 GRBs above 20 MeV, of which about 5 with the spark chamber at the higher energies (i.e. higher than 200 MeV), reflects broadly the LAT-vs-GBM statistics and indicates an intrinsic paucity of detected very high energy tails. Similar to the LAT-detected GRBs, the EGRET-detected ones were among the brightest BATSE GRBs (Dingus et al. 1995). Recently, Kaneko et al. (2008) analyzed the spectral shapes of the BATSE GRBs observed by the TASC calorimeter, i.e. with MeV emission, and found that the spectra of these bursts are quite hard (i.e. have fairly high E$_{\\rm peak}$ or small high energy spectral index). This is similar to what we find for the bursts detected by LAT (see Fig. 1). A proper understanding of the link between the GRB spectral maximum emission (100 keV - 1 MeV) and its countepart above few MeV is the key to get insight into the inner mechanism of power generation and into the jet formation and collimation (see Kumar \\& Barniol Duran 2009, Zhang \\& Pe'er 2009, Zou, Fan, Piran 2009, Li 2009), and ultimately will explain how the GRB emission at the highest energies correlates with the fundamental GRB parameters (Amati, Frontera, Guidorzi 2009, Ghirlanda, Nava, \\& Ghisellini 2009). In this paper, after having verified that the properties of the GBM GRB population do not differ from those of the previous GRB missions BATSE and Swift-BAT (Section 2), we analyze the sensitivity properties of the LAT and we estimate (following Band et al. 2009) the LAT detection rate of the high energy (100 MeV - 10 GeV) counterparts of GBM GRBs, under the hypothesis of a simple extrapolation of the GRB spectrum to energies larger than 1 MeV. \\begin{figure}[ht!] \\centerline{\\psfig{figure=spectrLAT.eps,height=8cm,width=9.cm}} \\caption{The spectral parameters, $\\beta$ and $E_{p}$ of the GRBs detected by LAT} \\end{figure} ", "conclusions": "We have shown that the simple extension of the synchrotron emission cannot justify the lack of detection of high energy emission from GRBs. The emission is suppressed at high energy by some mechanism (Fan 2009): i.e. the electrons are not accelerated to such an high energy, the high energy photons pair produce with low energy photons in the source and cannot escape the source or pair produce with external photons. On the other hand for the bursts detected by LAT there are two GRBs that have very soft synchrotron emission and their high energy emission cannot be explained by simple extrapolation of the low energy emission. There is evidence of another component at least in one GRB 090902b. \\begin{figure}[ht!] \\centerline{\\psfig{figure=beta.eps,height=8cm,width=9.cm}} \\caption{ The ratio s defined in the text as a function of the high energy spectral index $\\beta$ for each burst not detected by LAT (blue cross), and detected by LAT (red strars). The measured value of 080916C is also reported (black diamond)} \\end{figure} \\begin{figure}[ht!] \\centerline{\\psfig{figure=epeaks.eps,height=8cm,width=9.cm}} \\caption{ The ratio s defined in the text as a function of the peak energy for each burst not detected by LAT (blue cross), and detected by LAT (red strars). The measured value of 080916C is also reported (black diamond)} \\end{figure} \\onecolumn \\begin{table}[t] \\caption{GBM parameters for the bursts detected by Fermi. The first column report the GRB name, the second column the duration of the GRB, the third one the GRB fluence (set to -1 when not available), the fourth the peak flux (taken in 1 sec for long bursts and set to -1 when not available), the fifth column report the method used to fit the spectra, the sixth the peak energy (set to -1 when not available), the seventh and eight the $\\alpha$ and $\\beta$ spectral indexes (set to 1000 when not available). The ninth column report the angle, $\\theta$, from the LAT boresight, in deg (set to -1 when not available). The last column indicates LAT detection (1 = YES, 0 = NO)} \\begin{tabular}{lllllllllll} \\hline GRB & T90 & Fluence (erg/cm2) & PF(ph/s/cm2) & Function & $E_{peak}$ & $\\alpha$ & $\\beta$ & $\\theta$ & LAT \\\\ & sec & $10^{-6}$erg/cm$^2$& ph/cm$^2$/sec& & keV & & & \\\\ \\hline 080810 & 122.0 & 6.90e+00 & 1.85e+00 & PL+HEC & 313.5 & -0.91 & -1000.00 & 61 & 0 \\\\ 080812 & 15.0 & -1.00e+00 & -1.00e+00 & PL+HEC & 140.0 & 0.17 & -1000.00 & 71 & 0 \\\\ 080816A & 70.0 & 1.86e+01 & 3.48e+00 & PL+HEC & 146.7 & -0.57 & -1000.00 & 55 & 0 \\\\ 080816B & 5.0 & -1.00e+00 & 1.38e+00 & PL+HEC & 1230.0 & -0.37 & -1000.00 & 70 & 0 \\\\ 080817A & 70.0 & -1.00e+00 & -1.00e+00 & * & -1.0 & 1000.00 & -1000.00 & 80 & 0 \\\\ 080817B & 6.0 & 2.60e+00 & -1.00e+00 & SPL & -1.0 & -1.07 & -1000.00 & 68 & 0 \\\\ 080818A & 50.0 & 2.26e+00 & -1.00e+00 & SPL & -1.0 & -1.57 & -1000.00 & 68 & 0 \\\\ 080818B & 10.0 & 1.00e+00 & -1.00e+00 & PL+HEC & 80.0 & -1.30 & -1000.00 & 68 & 0 \\\\ 080823 & 46.0 & 4.10e+00 & -1.00e+00 & PL+HEC & 164.7 & -1.20 & -1000.00 & 77 & 0 \\\\ 080824 & 28.0 & 2.30e+00 & -1.00e+00 & Band & 100.0 & -0.40 & -2.10 & 17 & 0 \\\\ 080825C & 22.0 & 2.40e+01 & -1.00e+00 & Band & 155.0 & -0.39 & -2.34 & 60 & 1 \\\\ 080830 & 45.0 & 4.60e+00 & -1.00e+00 & Band & 154.0 & -0.59 & -1.69 & 23 & 0 \\\\ 080904 & 22.0 & 2.25e+00 & 3.50e+00 & Band & 35.0 & 0.00 & -2.70 & 23 & 0 \\\\ 080905A & 1.0 & 2.80e-01 & 6.10e+00 & SPL & -1.0 & -0.96 & -1000.00 & 28 & 0 \\\\ 080905B & 159.0 & 4.10e-02 & 2.10e-01 & SPL & -1.0 & -1.75 & -1000.00 & 82 & 0 \\\\ 080905C & 28.0 & 4.60e+00 & 4.40e+00 & PL+HEC & 78.8 & -0.90 & -1000.00 & 108 & 0 \\\\ 080906B & 5.0 & 1.09e+01 & 2.20e+01 & Band & 125.3 & -0.07 & -2.10 & 32 & 0 \\\\ 080912 & 17.0 & 3.30e+00 & 4.10e+00 & SPL & -1.0 & -1.74 & -1000.00 & 56 & 0 \\\\ 080913B & 140.0 & 2.20e+00 & -1.00e+00 & PL+HEC & 114.0 & -0.69 & -1000.00 & 71 & 0 \\\\ 080916A & 60.0 & 1.50e+01 & 4.50e+00 & PL+HEC & 109.0 & -0.90 & -1000.00 & 76 & 0 \\\\ 080916C & 100.9 & 2.40e+02 & 6.87e+00 & Band & 566.0 & -0.92 & -2.28 & 48 & 1 \\\\ 080920 & 85.0 & 2.40e+00 & 1.29e+00 & SPL & -1.0 & -1.42 & -1000.00 & 16 & 0 \\\\ 080925 & 29.0 & 9.70e+00 & -1.00e+00 & Band & 120.0 & -0.53 & -2.26 & 38 & 0 \\\\ 080927 & 25.0 & 5.70e+00 & 2.00e+00 & SPL & -1.0 & -1.50 & -1000.00 & 75 & 0 \\\\ 080928 & 87.0 & 1.50e+00 & -1.00e+00 & SPL & -1.0 & -1.80 & -1000.00 & -1 & 0 \\\\ 081003C & 67.0 & 5.40e+00 & -1.00e+00 & SPL & -1.0 & -1.41 & -1000.00 & 48 & 0 \\\\ 081006A & 7.0 & 7.10e-01 & -1.00e+00 & Band & 1135.0 & -0.77 & -1.80 & 16 & 0 \\\\ 081006B & 9.0 & 7.30e-01 & -1.00e+00 & SPL & -1.0 & -1.30 & -1000.00 & 3 & 0 \\\\ 081007A & 12.0 & 1.20e+00 & 2.20e+00 & SPL & -1.0 & -2.10 & -1000.00 & 116 & 0 \\\\ 081007B & 0.5 & -1.00e+00 & -1.00e+00 & * & -1.0 & 1000.00 & -1000.00 & -1 & 0 \\\\ 081009 & 9.0 & 3.50e+01 & -1.00e+00 & Band & 47.4 & 0.10 & -3.42 & 67 & 0 \\\\ 081012 & 30.0 & 3.80e+00 & 2.00e+00 & PL+HEC & 360.0 & -0.31 & -1000.00 & 61 & 0 \\\\ 081021 & 25.0 & 5.30e+00 & 4.20e+00 & Band & 117.0 & 0.11 & -2.80 & 125 & 0 \\\\ 081024B & 0.8 & 3.40e-01 & 4.20e+00 & PL+HEC & 1583.0 & -0.70 & -1000.00 & -1 & 1 \\\\ 081024C & 65.0 & 4.00e+00 & 1.00e+00 & Band & 65.0 & -0.60 & -2.50 & 78 & 0 \\\\ 081025 & 45.0 & 7.10e+00 & 4.50e+00 & Band & 200.0 & 0.15 & -2.05 & -1 & 0 \\\\ 081028B & 20.0 & 2.00e+00 & 6.90e+00 & PL+HEC & 70.0 & -0.55 & -1000.00 & 107 & 0 \\\\ 081101A & 0.2 & 1.60e-01 & -1.00e+00 & SPL & -1.0 & -1.14 & -1000.00 & -1 & 0 \\\\ 081101B & 8.0 & 1.60e+01 & 1.03e+01 & PL+HEC & 550.0 & -0.62 & -1000.00 & 116 & 0 \\\\ \\hline \\end{tabular} \\end{table} \\pagebreak \\newpage \\begin{table}[t] \\begin{tabular}{llllllllll} \\hline GRB name & t90 & fluence & peakflux & Function & E$_{\\rm peak}$ & $\\alpha$ & $\\beta$ & $\\theta$ & LAT\\\\ & sec & $10^{-6}$erg/cm$^2$& ph/cm$^2$/sec& keV & & & \\\\ \\hline 081102A & 88.0 & 2.10e+00 & -1.00e+00 & Band & 72.0 & 0.44 & -2.36 & -1 & 0 \\\\ 081102B & 2.2 & 1.12e+00 & 3.68e+00 & SPL & -1.0 & -1.07 & -1000.00 & 53 & 0 \\\\ 081105B & 0.2 & 2.28e-01 & 2.00e+01 & SPL & -1.0 & -1.17 & -1000.00 & 87 & 0 \\\\ 081107 & 2.2 & 1.64e+00 & 1.10e+01 & Band & 65.0 & 0.25 & -2.80 & 52 & 0 \\\\ 081109A & 45.0 & 6.53e+00 & 3.20e+00 & PL+HEC & 240.0 & -1.28 & -1000.00 & -1 & 0 \\\\ 081110 & 20.0 & -1.00e+00 & -1.00e+00 & * & -1.0 & 1000.00 & -1000.00 & 67 & 0 \\\\ 081113 & 0.5 & 1.07e+00 & 2.00e+01 & SPL & -1.0 & -1.28 & -1000.00 & 60 & 0 \\\\ 081118B & 20.0 & 1.12e-01 & 6.70e-01 & Band & 41.2 & 0.80 & -2.14 & 41 & 0 \\\\ 081119 & 0.8 & 4.10e-01 & 7.20e+00 & SPL & -1.0 & 1.30 & -1000.00 & 86 & 0 \\\\ 081120 & 12.0 & 2.70e+00 & 5.10e+00 & Band & 44.0 & 0.40 & -2.18 & 84 & 0 \\\\ 081121 & -1.0 & -1.00e+00 & -1.00e+00 & * & -1.0 & 1000.00 & -1000.00 & -1 & 0 \\\\ 081122A & 26.0 & 9.60e+00 & 3.00e+01 & Band & 158.6 & -0.63 & -2.24 & 21 & 0 \\\\ 081122B & 0.3 & 7.90e-02 & 1.60e+00 & SPL & -1.0 & -1.50 & -1000.00 & 52 & 0 \\\\ 081124 & 35.0 & 9.50e-02 & 6.70e-01 & Band & 22.8 & -0.60 & -2.83 & 86 & 0 \\\\ 081125 & 15.0 & 4.91e+01 & 2.70e+01 & Band & 221.0 & 0.14 & -2.34 & 126 & 0 \\\\ 081129 & 59.0 & 2.00e+01 & 1.40e+01 & Band & 150.0 & -0.50 & -1.84 & 118 & 0 \\\\ 081130A & -1.0 & -1.00e+00 & -1.00e+00 & * & -1.0 & 1000.00 & -1000.00 & 90 & 0 \\\\ 081130B & 12.0 & 1.30e+00 & 1.80e+00 & PL+HEC & 152.0 & -0.77 & -1000.00 & 66 & 0 \\\\ 081204B & 0.3 & 4.88e-01 & 1.63e+01 & SPL & -1.0 & -1.18 & -1000.00 & 46 & 0 \\\\ 081204C & 4.7 & 1.48e+00 & 7.20e+00 & SPL & -1.0 & -1.40 & -1000.00 & 56 & 0 \\\\ 081206A & 24.0 & 4.00e+00 & 2.40e+00 & Band & 151.0 & 0.13 & -2.20 & 102 & 0 \\\\ 081206B & 10.0 & -1.00e+00 & -1.00e+00 & * & -1.0 & 1000.00 & -1000.00 & 82 & 0 \\\\ 081206C & 20.0 & 1.19e+00 & 7.60e-01 & SPL & -1.0 & -1.35 & -1000.00 & 71 & 0 \\\\ 081207 & 153.0 & 1.06e+02 & -1.00e+00 & Band & 639.0 & -0.65 & -2.41 & 56 & 0 \\\\ 081209 & 0.4 & 5.90e-01 & 7.80e+00 & Band & 808.0 & -0.50 & -2.00 & 109 & 0 \\\\ 081213 & 0.0 & -1.00e+00 & -1.00e+00 & * & -1.0 & 1000.00 & -1000.00 & 51 & 0 \\\\ 081215A & 7.7 & 5.44e+01 & 6.89e+01 & Band & 304.0 & -0.58 & -2.07 & 86 & 1 \\\\ 081215B & 90.0 & 2.80e+00 & -1.00e+00 & PL+HEC & 139.0 & -0.14 & -1000.00 & 112 & 0 \\\\ 081216 & 1.0 & 3.60e+00 & 5.50e+01 & Band & 1235.0 & -0.70 & -2.17 & 95 & 0 \\\\ 081217 & 39.0 & 1.00e+01 & 4.00e+00 & Band & 167.0 & -0.61 & -2.70 & 54 & 0 \\\\ 081221 & 40.0 & 3.70e+01 & 3.30e+01 & Band & 77.0 & -0.42 & -2.91 & 78 & 0 \\\\ 081222 & 30.0 & 1.35e+01 & 1.48e+01 & Band & 134.0 & -0.55 & -2.10 & 50 & 0 \\\\ 081223 & 0.9 & 1.20e+00 & 2.20e+01 & PL+HEC & 280.0 & -0.63 & -1000.00 & 28 & 0 \\\\ 081224 & 50.0 & -1.00e+00 & -1.00e+00 & * & -1.0 & 1000.00 & -1000.00 & 16 & 0 \\\\ 081225 & 42.0 & 2.45e+00 & 6.00e-01 & SPL & -1.0 & -1.51 & -1000.00 & 55 & 0 \\\\ 081226A & 1.7 & 2.10e-01 & 3.20e+00 & SPL & -1.0 & 1.17 & -1000.00 & 110 & 0 \\\\ 081226B & 0.4 & 6.10e-01 & 1.77e+01 & Band & 300.0 & -0.20 & -1.82 & 22 & 0 \\\\ 081226C & 60.0 & 2.32e+00 & 4.50e+00 & PL+HEC & 82.0 & -1.04 & -1000.00 & 54 & 0 \\\\ 081229 & 0.5 & 8.70e-01 & 1.07e+01 & Band & 585.0 & -0.27 & -2.00 & 44 & 0 \\\\ \\hline \\end{tabular} \\end{table} \\pagebreak \\newpage \\begin{table}[t] \\begin{tabular}{llllllllll} \\hline GRB name & t90 & fluence & peakflux & Function & E$_{\\rm peak}$ & $\\alpha$ & $\\beta$ & $\\theta$ & LAT\\\\ & sec & $10^{-6}$erg/cm$^2$& ph/cm$^2$/sec& keV & & & \\\\ \\hline 081231 & 29.0 & 1.20e+01 & 1.53e+00 & Band & 152.3 & -0.80 & -2.03 & 20 & 0 \\\\ 090107B & 24.1 & 1.75e+00 & 3.68e+00 & PL+HEC & 106.1 & -0.68 & -1000.00 & -1 & 0 \\\\ 090108A & 0.9 & 1.28e+00 & 3.97e+01 & Band & 104.8 & -0.47 & -1.97 & 60 & 0 \\\\ 090108B & 0.8 & 7.90e-01 & 1.90e+01 & SPL & -1.0 & -0.99 & -1000.00 & 72 & 0 \\\\ 090109 & 5.0 & 1.21e+00 & 2.76e+00 & SPL & -1.0 & -1.50 & -1000.00 & 62 & 0 \\\\ 090112A & 65.0 & 5.20e+00 & 7.00e+00 & Band & 150.0 & -0.94 & -2.01 & 4 & 0 \\\\ 090112B & 12.0 & 5.40e+00 & 1.40e+01 & Band & 139.0 & -0.75 & -2.43 & 95 & 0 \\\\ 090117A & 21.0 & 1.80e+00 & 9.60e+00 & Band & 25.0 & -0.40 & -2.50 & 51 & 0 \\\\ 090117B & 27.0 & 2.10e+00 & 4.60e+00 & SPL & -1.0 & -1.55 & -1000.00 & 49 & 0 \\\\ 090117C & 86.0 & 1.10e+01 & 4.20e+00 & Band & 247.0 & -1.00 & -2.10 & 54 & 0 \\\\ 090126B & 10.8 & 1.25e+00 & 4.90e+00 & PL+HEC & 47.5 & -0.99 & -1000.00 & 18 & 0 \\\\ 090126C & -1.0 & -1.00e+00 & -1.00e+00 & * & -1.0 & 1000.00 & -1000.00 & 68 & 0 \\\\ 090129 & 17.2 & 5.60e+00 & 8.00e+00 & Band & 123.2 & -1.39 & -1.98 & 22 & 0 \\\\ 090131 & 36.4 & 2.23e+01 & 4.79e+01 & Band & 58.4 & -1.27 & -2.26 & 40 & 0 \\\\ 090202 & 66.0 & 8.65e+00 & 7.77e+00 & PL+HEC & 570.0 & -1.31 & -1000.00 & 55 & 0 \\\\ 090206 & 0.8 & 1.04e+00 & 1.90e+01 & PL+HEC & 710.0 & -0.65 & -1000.00 & 72 & 0 \\\\ 090207 & 10.0 & 4.01e+00 & 1.88e+00 & SPL & -1.0 & -1.59 & -1000.00 & 45 & 0 \\\\ 090213 & -1.0 & -1.00e+00 & -1.00e+00 & * & -1.0 & 1000.00 & -1000.00 & 17 & 0 \\\\ 090217 & 32.8 & 3.08e+01 & 1.12e+01 & Band & 610.0 & -0.85 & -2.86 & 34 & 1 \\\\ 090219 & 0.5 & 8.00e-01 & 7.70e+00 & SPL & -1.0 & -1.43 & -1000.00 & 137 & 0 \\\\ 090222 & 18.0 & 2.19e+00 & 1.10e+00 & Band & 147.9 & -0.97 & -2.56 & 80 & 0 \\\\ 090227A & 50.0 & 9.00e+00 & 4.57e+00 & Band & 1357.0 & -0.92 & -3.60 & 21 & 0 \\\\ 090227B & 0.9 & 8.70e+00 & 3.46e+01 & Band & 2255.0 & -0.53 & -3.04 & 72 & 0 \\\\ 090228A & 0.8 & 6.10e+00 & 1.33e+02 & Band & 849.0 & -0.35 & -2.98 & 16 & 0 \\\\ 090228B & 7.2 & 9.96e-01 & 2.53e+00 & PL+HEC & 147.8 & -0.70 & -1000.00 & 20 & 0 \\\\ 090301B & 28.0 & 2.69e+00 & 4.40e+00 & Band & 427.0 & -0.98 & -1.93 & 56 & 0 \\\\ 090303 & -1.0 & -1.00e+00 & -1.00e+00 & * & -1.0 & 1000.00 & -1000.00 & -1 & 0 \\\\ 090304 & -1.0 & -1.00e+00 & -1.00e+00 & * & -1.0 & 1000.00 & -1000.00 & 31 & 0 \\\\ 090305B & 2.0 & 2.70e+00 & 1.10e+01 & Band & 770.0 & -0.50 & -1.90 & 40 & 0 \\\\ 090306C & 38.8 & 9.00e-01 & 2.40e+00 & Band & 87.0 & -0.32 & -2.28 & 14 & 0 \\\\ 090307B & 30.0 & 1.70e+00 & 1.80e+00 & PL+HEC & 212.0 & -0.70 & -1000.00 & 83 & 0 \\\\ 090308B & 2.1 & 3.46e+00 & 1.42e+01 & PL+HEC & 710.3 & -0.54 & -1000.00 & 50 & 0 \\\\ 090309B & 60.0 & 4.70e+00 & 4.43e+00 & PL+HEC & 197.0 & -1.52 & -1000.00 & 26 & 0 \\\\ 090310 & 125.2 & 2.15e+00 & 4.40e+00 & PL+HEC & 279.0 & -0.65 & -1000.00 & 77 & 0 \\\\ 090319 & 67.7 & 7.47e+00 & 3.85e+00 & PL+HEC & 187.3 & 0.90 & -1000.00 & 27 & 0 \\\\ 090320A & 10.0 & -1.00e+00 & -1.00e+00 & * & -1.0 & 1000.00 & -1000.00 & 60 & 0 \\\\ 090320B & 52.0 & 1.10e+00 & 1.20e-01 & PL+HEC & 72.0 & -1.10 & -1000.00 & 101 & 0 \\\\ 090320C & 4.0 & -1.00e+00 & -1.00e+00 & * & -1.0 & 1000.00 & -1000.00 & 40 & 0 \\\\ 090323 & 70.0 & 1.00e+02 & 1.23e+01 & PL+HEC & 697.0 & -0.89 & -1000.00 & -1 & 1 \\\\ \\hline \\end{tabular} \\end{table} \\pagebreak \\newpage \\begin{table}[t] \\begin{tabular}{llllllllll} \\hline GRB name & t90 & fluence & peakflux & Function & E$_{\\rm peak}$ & $\\alpha$ & $\\beta$ & $\\theta$ & LAT\\\\ & sec & $10^{-6}$erg/cm$^2$& ph/cm$^2$/sec& keV & & & \\\\ \\hline 090326 & 11.2 & 8.60e-01 & -1.00e+00 & PL+HEC & 75.0 & -0.86 & -1000.00 & 103 & 0 \\\\ 090327 & 24.0 & 3.00e+00 & 3.50e+00 & Band & 89.7 & -0.39 & -2.90 & 66 & 0 \\\\ 090328A & 100.0 & 8.09e+01 & 1.85e+01 & Band & 653.0 & -0.93 & -2.20 & -1 & 1 \\\\ 090328B & 0.3 & 9.61e-01 & 2.98e+01 & Band & 1967.0 & -0.92 & -2.48 & 74 & 0 \\\\ 090330 & 80.0 & 1.14e+01 & 6.80e+00 & Band & 246.0 & -0.99 & -2.68 & 50 & 0 \\\\ 090331 & -1.0 & -1.00e+00 & -1.00e+00 & * & -1.0 & 1000.00 & -1000.00 & 40 & 0 \\\\ 090403 & 16.0 & -1.00e+00 & -1.00e+00 & * & -1.0 & 1000.00 & -1000.00 & 42 & 0 \\\\ 090405 & 1.2 & -1.00e+00 & -1.00e+00 & * & -1.0 & 1000.00 & -1000.00 & 70 & 0 \\\\ 090409 & 20.0 & 6.14e-01 & 1.36e+00 & PL+HEC & 137.0 & 1.20 & -1000.00 & 90 & 0 \\\\ 090411A & 24.6 & 8.60e+00 & 3.25e+00 & Band & 141.0 & -0.88 & -1.82 & 59 & 0 \\\\ 090411B & 18.7 & 8.00e+00 & 7.40e+00 & Band & 189.0 & -0.80 & -2.00 & 111 & 0 \\\\ 090412 & 0.5 & -1.00e+00 & -1.00e+00 & * & -1.0 & 1000.00 & -1000.00 & 71 & 0 \\\\ 090418C & 0.6 & 6.00e-01 & 8.50e+00 & PL+HEC & 1000.0 & -0.94 & -1000.00 & 58 & 0 \\\\ 090422 & 10.0 & 1.00e+00 & 7.80e+00 & SPL & -1.0 & 1.81 & -1000.00 & 29 & 0 \\\\ 090423 & 12.0 & 1.10e+00 & 3.30e+00 & PL+HEC & 82.0 & -0.77 & -1000.00 & 75.6 & 0 \\\\ 090424 & 52.0 & 5.20e+01 & 1.37e+02 & Band & 177.0 & 0.90 & -2.90 & 71 & 0 \\\\ 090425 & 72.0 & 1.30e+01 & 1.40e+01 & Band & 69.0 & -1.29 & -2.03 & 105 & 0 \\\\ 090426B & 3.8 & 5.20e-01 & -1.00e+00 & SPL & -1.0 & -1.60 & -1000.00 & 56 & 0 \\\\ 090426C & 12.0 & 3.10e+00 & 6.80e+00 & Band & 295.0 & -1.29 & -1.98 & 69 & 0 \\\\ 090427B & 7.0 & 8.00e-01 & -1.00e+00 & SPL & -1.0 & -1.10 & -1000.00 & 14 & 0 \\\\ 090427C & 12.5 & 1.60e+00 & -1.00e+00 & PL+HEC & 75.0 & 0.35 & -1000.00 & 81 & 0 \\\\ 090428A & 8.0 & 9.90e-01 & 1.23e+01 & Band & 85.0 & -0.40 & -2.70 & 96 & 0 \\\\ 090428B & 30.0 & 5.20e+00 & 1.01e+01 & Band & 53.0 & -1.81 & -2.17 & 101 & 0 \\\\ 090429C & 13.0 & 3.70e+00 & 6.70e+00 & SPL & -1.0 & -1.43 & -1000.00 & 112 & 0 \\\\ 090429D & 11.0 & 1.60e+00 & 8.60e-01 & SPL & 223.0 & -0.87 & -1000.00 & 33 & 0 \\\\ 090502 & 66.2 & 3.50e-02 & 6.20e+00 & PL+HEC & 63.2 & -1.10 & -1000.00 & 77 & 0 \\\\ 090509 & 295.0 & 8.40e+00 & 3.10e+00 & PL+HEC & 343.0 & -0.90 & -1000.00 & 75 & 0 \\\\ 090510A & 1.4 & 3.00e+01 & 8.00e+01 & Band & 4400.0 & -0.80 & -2.60 & -1 & 1 \\\\ 090510B & 7.0 & -1.00e+00 & -1.00e+00 & * & -1.0 & 1000.00 & -1000.00 & 100 & 0 \\\\ 090511 & 14.0 & 1.80e+00 & 2.50e+00 & PL+HEC & 391.0 & -0.95 & -1000.00 & 67 & 0 \\\\ 090513A & 23.0 & 6.80e+00 & 2.70e+00 & PL+HEC & 850.0 & -0.90 & -1000.00 & 89 & 0 \\\\ 090513B & -1.0 & -1.00e+00 & -1.00e+00 & * & -1.0 & 1000.00 & -1000.00 & 119 & 0 \\\\ 090514 & 49.0 & 8.10e+00 & 7.60e+00 & SPL & -1.0 & -1.92 & -1000.00 & 19 & 0 \\\\ 090516A & 350.0 & 2.30e+01 & 5.30e+00 & Band & 51.4 & -1.03 & -2.10 & 20 & 0 \\\\ 090516B & 350.0 & 3.00e+01 & 4.00e+00 & PL+HEC & 327.0 & -1.01 & -1000.00 & 45 & 0 \\\\ 090516C & 15.0 & 4.00e+00 & 7.70e+00 & Band & 38.0 & -0.44 & -1.81 & 69 & 0 \\\\ 090518A & 9.0 & 1.60e+00 & 4.70e+00 & SPL & -1.0 & -1.59 & -1000.00 & 53 & 0 \\\\ 090518B & 12.0 & 2.20e+00 & 5.60e+00 & PL+HEC & 127.0 & -0.74 & -1000.00 & 90 & 0 \\\\ 090519B & 87.0 & 1.40e+00 & 5.02e+00 & SPL & -1.0 & -1.63 & -1000.00 & 18 & 0 \\\\ 090520B & 1.5 & 4.50e-01 & 4.10e+00 & SPL & -1.0 & -1.40 & -1000.00 & 10 & 0 \\\\ 090520C & 4.9 & 3.54e+00 & 4.47e+00 & Band & 204.2 & -0.73 & -1.96 & 71 & 0 \\\\ 090520D & 12.0 & 4.00e+00 & 4.10e+00 & Band & 46.3 & -0.99 & -3.25 & 66 & 0 \\\\ 090522 & 22.0 & 1.20e+00 & 3.50e+00 & PL+HEC & 75.8 & -1.03 & -1000.00 & 53 & 0 \\\\ 090524 & 72.0 & 1.85e+01 & 1.41e+01 & Band & 82.6 & -1.00 & -2.30 & 63 & 0 \\\\ 090528A & 68.0 & 9.30e+00 & 7.60e+00 & PL+HEC & 99.0 & -1.70 & -1000.00 & 81 & 0 \\\\ 090528B & 102.0 & 4.65e+01 & 1.47e+01 & Band & 172.0 & -1.10 & -2.30 & 65 & 0 \\\\ \\hline \\end{tabular} \\end{table} \\pagebreak \\newpage \\begin{table}[t] \\begin{tabular}{llllllllll} \\hline GRB name & t90 & fluence & peakflux & Function & E$_{\\rm peak}$ & $\\alpha$ & $\\beta$ & $\\theta$ & LAT\\\\ & sec & $10^{-6}$erg/cm$^2$& ph/cm$^2$/sec& keV & & & \\\\ \\hline 090529B & 5.1 & 3.40e-01 & 4.10e+00 & Band & 142.0 & -0.70 & -2.00 & 36 & 0 \\\\ 090529C & 10.4 & 3.10e+00 & 2.50e+01 & Band & 188.0 & -0.84 & -2.10 & 69 & 0 \\\\ 090530B & 194.0 & 5.90e+01 & 1.08e+01 & Band & 67.0 & -0.71 & -2.42 & 84 & 0 \\\\ 090531B & 2.0 & 6.20e-01 & 1.49e+00 & Band & 2166.0 & -0.71 & -2.47 & 25 & 0 \\\\ 090602 & 16.0 & 5.70e+00 & 3.62e+00 & PL+HEC & 503.0 & -0.56 & -1000.00 & 112 & 0 \\\\ 090606 & 60.0 & 3.19e+00 & 2.41e+00 & SPL & -1.0 & -1.63 & -1000.00 & 128 & 0 \\\\ 090608 & 61.0 & 3.20e+00 & 2.70e+00 & SPL & -1.0 & -1.83 & -1000.00 & 93 & 0 \\\\ 090610A & 6.5 & 7.32e-01 & 9.40e-01 & SPL & -1.0 & -1.30 & -1000.00 & 70 & 0 \\\\ 090610B & 202.5 & 4.13e+00 & 1.54e+00 & PL+HEC & 104.9 & -0.46 & -1000.00 & 91 & 0 \\\\ 090610C & 18.1 & 8.54e-01 & 1.12e+00 & SPL & -1.0 & -1.62 & -1000.00 & 104 & 0 \\\\ 090612 & 58.0 & 2.37e+00 & 1.63e+00 & Band & 357.0 & -0.60 & -1.90 & 56 & 0 \\\\ 090616 & 2.7 & 2.23e-01 & 2.08e+00 & SPL & -1.0 & -1.27 & -1000.00 & 68 & 0 \\\\ 090617 & 0.4 & 4.68e-01 & 1.00e+01 & Band & 684.0 & -0.45 & -2.00 & 45 & 0 \\\\ 090618 & 155.0 & 2.70e+02 & 7.34e+01 & Band & 155.5 & -1.26 & -2.50 & 133 & 0 \\\\ 090620 & 16.5 & 6.60e+00 & 7.00e+00 & Band & 156.0 & -0.40 & -2.44 & 60 & 0 \\\\ 090621A & 294.0 & 4.40e+00 & 1.92e+00 & Band & 56.0 & -1.10 & -2.12 & 12 & 0 \\\\ 090621B & 0.1 & 3.71e-01 & 6.40e+00 & Band & 321.6 & -0.13 & -1.57 & 108 & 0 \\\\ 090621C & 59.9 & 1.80e+00 & 2.29e+00 & PL+HEC & 148.0 & -1.40 & -1000.00 & 52 & 0 \\\\ 090621D & 39.9 & 1.34e+00 & 1.74e+00 & SPL & -1.0 & -1.66 & -1000.00 & 79 & 0 \\\\ 090623 & 72.2 & 9.60e+00 & 3.30e+00 & Band & 428.0 & -0.69 & -2.30 & 73 & 0 \\\\ 090625A & 51.0 & 8.80e-01 & 5.00e-01 & PL+HEC & 198.0 & -0.60 & -1000.00 & 13 & 0 \\\\ 090625B & 13.6 & 1.04e+00 & 1.87e+00 & Band & 100.0 & -0.40 & -2.00 & 125 & 0 \\\\ 090626 & 70.0 & 3.50e+01 & 1.79e+01 & Band & 175.0 & -1.29 & -1.98 & 15 & 1 \\\\ 090630 & 5.1 & 5.10e-01 & 2.78e+00 & Band & 71.0 & -1.50 & -2.30 & 75 & 0 \\\\ 090701 & 12.0 & 4.50e-01 & 2.10e+00 & SPL & -1.0 & 1.84 & -1000.00 & 13 & 0 \\\\ 090703 & 9.0 & 6.80e-01 & 1.00e+00 & SPL & -1.0 & -1.72 & -1000.00 & 25 & 0 \\\\ 090704 & 70.0 & 5.80e+00 & 1.20e+00 & PL+HEC & 233.7 & -1.13 & -1000.00 & 77 & 0 \\\\ 090706 & 100.0 & 1.50e+00 & 1.24e+00 & SPL & -1.0 & -2.16 & -1000.00 & 20 & 0 \\\\ 090708 & 18.0 & 4.00e-01 & 1.00e+00 & PL+HEC & 47.5 & -1.29 & -1000.00 & 55 & 0 \\\\ 090709B & 32.0 & 1.30e+00 & 2.00e+00 & PL+HEC & 130.0 & -1.01 & -1000.00 & 35 & 0 \\\\ 090711 & 100.0 & 1.17e+01 & 4.20e+00 & PL+HEC & 210.0 & -1.30 & -1000.00 & 13 & 0 \\\\ 090712 & 72.0 & 4.20e+00 & 6.30e-01 & PL+HEC & 505.0 & -0.68 & -1000.00 & 33 & 0 \\\\ 090713 & 113.0 & 3.70e+00 & 1.60e+00 & PL+HEC & 99.0 & -0.34 & -1000.00 & 63 & 0 \\\\ 090717A & 70.0 & 4.50e-01 & 7.80e+00 & Band & 120.0 & -0.88 & -2.33 & 70 & 0 \\\\ 090717B & 0.9 & 4.83e-01 & 3.91e+00 & SPL & -1.0 & -1.02 & -1000.00 & 35 & 0 \\\\ 090718A & -1.0 & -1.00e+00 & -1.00e+00 & * & -1.0 & 1000.00 & -1000.00 & 51 & 0 \\\\ 090718B & 28.0 & 2.52e+01 & 3.20e+01 & Band & 184.0 & -1.18 & -2.59 & 76 & 0 \\\\ \\hline \\end{tabular} \\end{table} \\pagebreak \\newpage \\begin{table}[t] \\begin{tabular}{llllllllll} \\hline GRB name & t90 & fluence & peakflux & Function & E$_{\\rm peak}$ & $\\alpha$ & $\\beta$ & $\\theta$ & LAT\\\\ & sec & $10^{-6}$erg/cm$^2$& ph/cm$^2$/sec& keV & & & \\\\ \\hline 090719 & 16.0 & 4.83e+01 & 3.78e+01 & Band & 254.0 & -0.68 & -2.92 & 88 & 0 \\\\ 090720A & 7.0 & 2.90e+00 & 1.09e+01 & PL+HEC & 117.5 & -0.75 & -1000.00 & 113 & 0 \\\\ 090720B & 20.0 & 1.06e+01 & 1.09e+01 & Band & 924.0 & -1.00 & -2.43 & 56 & 0 \\\\ 090802A & 0.1 & 6.50e-01 & 6.12e+01 & Band & 283.0 & -0.42 & -2.40 & 123 & 0 \\\\ 090802B & -1.0 & -1.00e+00 & -1.00e+00 & * & -1.0 & 1000.00 & -1000.00 & 104 & 0 \\\\ 090807B & 3.0 & 1.02e+00 & 1.09e+01 & Band & 37.0 & -0.60 & -2.40 & 45 & 0 \\\\ 090809B & 15.0 & 2.26e+01 & 2.36e+01 & Band & 198.0 & -0.85 & -2.02 & 81 & 0 \\\\ 090813 & 9.0 & 3.50e+00 & 1.44e+01 & Band & 95.0 & -1.25 & -2.00 & 35.3 & 0 \\\\ 090814C & 0.2 & 6.60e-01 & 9.10e+00 & PL+HEC & 790.0 & -0.39 & -1000.00 & 61 & 0 \\\\ 090815A & 200.0 & 3.40e+00 & 1.90e+00 & SPL & -1.0 & -1.50 & -1000.00 & 87 & 0 \\\\ 090815B & 30.0 & 5.05e+00 & 1.44e+01 & Band & 18.6 & -1.82 & -2.70 & 82 & 0 \\\\ 090817 & 220.0 & 7.30e+00 & 3.80e+00 & Band & 115.0 & -1.10 & -2.20 & 82 & 0 \\\\ 090820A & 60.0 & 6.60e+01 & 5.80e+01 & Band & 215.0 & -0.69 & -2.61 & 108 & 0 \\\\ 090820B & 11.2 & 1.16e+00 & 6.10e+00 & PL+HEC & 38.8 & -1.44 & -1000.00 & 32 & 0 \\\\ 090826 & 8.5 & 1.26e+00 & 3.28e+00 & PL+HEC & 172.0 & -0.96 & -1000.00 & 35 & 0 \\\\ 090828 & 100.0 & 2.52e+01 & 1.62e+01 & Band & 136.5 & -1.23 & -2.12 & 95 & 0 \\\\ 090829A & 85.0 & 1.02e+02 & 5.15e+01 & Band & 183.0 & -1.44 & -2.10 & 47 & 0 \\\\ 090829B & 100.0 & 6.40e+00 & 3.20e+00 & Band & 143.0 & -0.70 & -2.40 & 42 & 0 \\\\ 090831 & 69.1 & 1.66e+01 & 9.40e+00 & Band & 243.8 & -1.52 & -1.96 & 107 & 0 \\\\ 090902A & 1.2 & 2.11e+00 & 1.14e+01 & Band & 388.0 & 0.30 & -2.05 & 82 & 0 \\\\ 090902B & 21.0 & 3.74e+02 & 4.61e+01 & Band & 798.0 & -0.61 & -3.87 & 52 & 1 \\\\ 090904B & 71.0 & 2.44e+01 & 9.80e+00 & Band & 106.3 & -1.26 & -2.18 & 113 & 0 \\\\ 090910 & 62.0 & 9.20e+00 & 2.30e+00 & Band & 274.8 & -0.90 & -2.00 & 107 & 0 \\\\ 090922A & 92.0 & 1.14e+01 & 1.56e+01 & Band & 139.3 & -0.77 & -2.28 & 19 & 0 \\\\ 090925 & 50.0 & 9.46e+00 & 4.20e+00 & Band & 156.0 & -0.60 & -1.91 & 116 & 0 \\\\ 090926A & 20.0 & 1.45e+02 & 8.08e+01 & Band & 268.0 & -0.69 & -2.34 & 52 & 1 \\\\ 090926B & 81.0 & 8.70e+00 & -1.00e+00 & PL+HEC & 91.0 & -0.13 & -1000.00 & 100 & 0 \\\\ 090927 & 2.0 & 6.10e-01 & 7.20e+00 & SPL & -1.0 & -1.47 & -1000.00 & 85 & 0 \\\\ 090929A & 8.5 & 1.06e+01 & 1.09e+01 & PL+HEC & 610.9 & -0.52 & -1000.00 & 122 & 0 \\\\ 091003A & 21.1 & 3.76e+01 & 3.18e+01 & Band & 486.2 & -1.13 & -2.64 & 13 & 1 \\\\ \\label{t:fit} \\end{tabular} \\end{table} \\begin{table}[t] \\caption{ Best fit parameters $Rate(z=0)$ , $L^*$, $\\alpha$ and $\\beta$ and their 1-$\\sigma$ confidence levels. For each fit we report the $\\chi^2$ values corresponding to the best fit ($\\chi^2_{\\rm b.f.}$). Also shown are the parameter for a LF, LFb, that fit quite well all the samples, the $\\chi^2\\sim 1.3 $ for all the instruments } \\begin{tabular}{|c|c|c|c|c|c|} \\hline sample & Rate(z=0) & $L^*$ & $\\alpha$ & $\\beta$ &$\\chi^2$\\\\ & $Gpc^{-3} yr^{-1}$ & $10^{51}$ erg/sec & & & \\\\ \\hline GBM & $0.5^{+0.3}_{-0.2}$ & $5.5^{+1.5}_{-2}$ & $0.3^{+0.1}_{-0.5}$ &$2.3^{+0.6}_{-0.3}$ & 1.1 \\\\ BATSE & $1.0^{+0.2}_{-0.4}$ & $4_{-1.5}^{+2}$ & $0.1_{-0.05}^{+0.3}$ &$2.6_{-0.5}^{+0.9}$ & 1.1 \\\\ {\\it Swift} & $0.6^{+0.3}_{-0.1}$ & $3.3_{-0.5}^{+2.5}$ & $0.1_{-0.05}^{+0.3}$ &$2.7_{-0.4}^{+1}$ & 0.95 \\\\ LFb & $0.8$ & $5 $ & $0.1$ & $2.7$ & 1.3 \\\\ \\hline \\end{tabular} \\end{table}" }, "0910/0910.0241_arXiv.txt": { "abstract": "We generalize the Swiss-cheese cosmologies so as to include nonzero linear momenta of the associated boundary surfaces. The evolution of mass scales in these generalized cosmologies is studied for a variety of models for the background without having to specify any details within the local inhomogeneities. We find that the final effective gravitational mass and size of the evolving inhomogeneities depends on their linear momenta but these properties are essentially unaffected by the details of the background model. ", "introduction": "The Swiss-cheese models give us (noninteracting) inhomogeneities in a cosmological setting that are exact solutions to Einstein's equations. As a result, the models have become a standard construction \\cite{forms} and are very widely studied \\cite{2}. The classical Einstein-Straus vacuole model, which requires a comoving boundary surface, is unstable. Many subsequent Swiss-cheese models have also assumed that the associated boundary surfaces remain comoving, but it is well known that this need not be the case. Some general studies of this issue go back many years \\cite{lake}. Some specific examples of noncomoving boundary surfaces include the study of Vaidya-type inhomogeneities \\cite{fayos} and the evolution of ``density waves\" in the Lema\\^{\\i}tre-Tolman model \\cite{lt}. Here we examine the role of the linear momentum of a boundary surface in a Robertson-Walker background. Within the context of these generalized models, we study the evolution of the effective gravitational mass and size of the inhomogeneities for a variety of well-known background models. We find that whereas the momentum plays a central role, the details of the background model are relatively unimportant. ", "conclusions": "We have introduced a new generalized Swiss-cheese model which does not assume \\emph{a priori} that the associated boundary surfaces are comoving. In order to quantify evolving inhomogeneities, we have considered geodesic boundaries characterized by their linear momentum $\\mathcal{D}$. For the size of the inhomogeneities we are interested in, the physical values of $\\mathcal{D}/c$ are $\\sim10^{-5}$. For a given linear momentum, we have found that the inhomogeneities grow almost independently of the background model (with the inclusion of the radiation density parameter $\\Omega_{R_0}$). As shown in Fig.\\ref{rsigmabar} these inhomogeneities are almost at their full size by the decoupling ($z\\sim 1100$). For a redshift of $z \\lesssim 2$, corresponding to high redshift supernovae, the inhomogeneities considered here are growing very slowly as is shown in see Fig.\\ref{MSMLCDM}." }, "0910/0910.0077_arXiv.txt": { "abstract": "{ We present a large sample of candidate galaxies at $z\\approx 7$ -- 10, selected in the Hubble Ultra Deep Field using the new observations of the Wide Field Camera 3 that was recently installed to Hubble Space Telescope. Our sample is composed of 20 $z_{850}$-dropouts (four new discoveries), 15 $Y_{105}$-dropouts (nine new discoveries) and 20 $J_{125}$-dropouts (all new discoveries). The surface densities of the $z_{850}$-dropouts are close to what predicted by earlier studies, however, those of the $Y_{105}$- and $J_{125}$-dropouts are quite unexpected. While no $Y_{105}$- or $J_{125}$-dropouts have been found at $AB\\leq 28.0$~mag, their surface densities seem to increase sharply at fainter levels. While some of these candidates seem to be close to foreground galaxies and thus could possibly be gravitationally lensed, the overall surface densities after excluding such cases are still much higher than what would be expected if the luminosity function does not evolve from $z\\sim 7$ to 10. Motivated by such steep increases, we tentatively propose a set of Schechter function parameters to describe the luminosity functions at $z\\approx 8$ and 10. As compared to their counterpart at $z\\approx 7$, here $L^*$ decreases by a factor of $\\sim 6.5$ and $\\Phi^*$ increases by a factor of 17--90. Although such parameters are not yet demanded by the existing observations, they are allowed and seem to agree with the data better than other alternatives. If these luminosity functions are still valid beyond our current detection limit, this would imply a sudden emergence of a large number of low-luminosity galaxies when looking back in time to $z\\approx 10$, which, while seemingly exotic, would naturally fit in the picture of the cosmic hydrogen reionization. These early galaxies could easily account for the ionizing photon budget required by the reionization, and they would imply that the global star formation rate density might start from a very high value at $z\\approx 10$, rapidly reach the minimum at $z\\approx 7$, and start to rise again towards $z\\approx 6$. In this scenario, the majority of the stellar mass that the universe assembled through the reionization epoch seems still undetected by current observations at $z\\approx 6$. ", "introduction": "An important subject in cosmology is the hydrogen reionization of the universe. The detections of complete Gunn-Peterson troughs (Gunn \\& Peterson 1965) in the spectra of a few $z>6$ QSOs (see Fan et al. 2006) from the Sloan Digital Sky Survey indicate that the reionization must have ended at around $z\\approx 6$, while the recent measurement of Thomson scattering optical depth from the seven-year Wilkinson Microwave Anisotropy Probe (WMAP) data shows that the reionization most likely began at $z=10.4\\pm1.2$ if assuming an instantaneous reionization history (Komatsu et al. 2010). An important and yet controversial question is the sources of reionization. It is clear that the QSO population can only produce a small fraction of the necessary ionizing photons at $z\\approx 6$ (e.g., Fan et al. 2002), which leaves star-forming galaxies as the most obvious alternative, although the measurement of the star-formation rate density (and hence the ionizing photon budget) is still uncertain (e.g., Stanway et al. 2003; Bouwens et al. 2003; Giavalisco et al. 2004b; Bunker et al. 2004; Yan \\& Windhorst 2004b). Yan \\& Windhorst (2004a; hereafter YW04a) pointed out that star-forming galaxies could account for the entire ionizing photon budget at $z\\approx 6$ as long as their luminosity function (LF) has a steep faint-end slope, $\\alpha<-1.6$. Using the Hubble Ultra Deep Field (HUDF; Beckwith 2006) data obtained by the Advanced Camera for Surveys (ACS), Yan \\& Windhorst (2004b; hereafter YW04b) have found a large number of $i_{775}$-dropouts, which are candidate galaxies at $z\\approx 6$, and obtained their LF that indeed has $\\alpha\\lesssim -1.8$. Such a very steep slope was later confirmed by Bouwens et al. (2006) using a larger $i_{775}$-dropout sample collected in the HUDF, the HUDF parallel fields and the two fields of the Great Observatories Origins Deep Survey (GOODS; Giavalisco et al. 2004a). Bouwens et al. (2007; hereafter B07) further compared the LF at $z\\approx 4$, 5 and 6, and suggested that LF has a strong luminosity evolution over this period in that $M^*$ is $\\sim 0.7$~mag fainter at $z\\approx 6$ than at $z\\approx 4$. On the other hand, they found that the LF does not change much in $\\alpha$. As star-forming galaxies seem to be capable of keeping the universe ionized at $z\\approx 6$, it is natural to expect that star-forming galaxies are also capable of producing sufficient ionizing photons at earlier epochs to make the reionization happen, and therefore we should expect that they must exist in significant number extending well into the reionization epoch. Along a different line of study, it has also been concluded that the universe must have started actively forming stars long before $z\\approx 6$. Using the GOODS {\\it Spitzer} Infrared Array Camera (IRAC) data in the HUDF region, Yan et al. (2005) have detected the restframe optical fluxes from three $z\\approx 6$ and eleven $z\\approx 5$ galaxies in the 3.6 and 4.5~$\\mu$m channels, and found that these are rather matured systems with stellar masses to the order of $\\sim 10^{10}M_\\odot$ and ages to the order of a few hundred Myr (see also Eyles et al. 2005). This strongly suggests that such high-mass galaxies must have started forming their stars at $z>7$ and likely earlier. Yan et al. (2006) further investigate this problem using a much larger sample in the entire GOODS field, and reinforced this conclusion (see also Eyles et al. 2007; Stark et al. 2007). Therefore, one should indeed expect a significant number of galaxies at $z\\gtrsim 7$. To search for galaxies at $z\\gtrsim 7$, we have to carry out surveys in the near-IR regime, because the line-of-sight neutral hydrogen absorption effectively extincts the light from such sources that is emitted below 1~$\\mu m$ in observer's frame. In fact, we rely on this effect to identify such galaxies. The first candidate of this kind was reported by Dickinson et al. (2000) using the data obtained by the Near Infrared Camera and Multi-Object Spectrometer (NICMOS) No. 3 (NIC3) in the Hubble Deep Field, although now we have convincing evidence that it is likely a red galaxy at lower redshift, similar to those ``IRAC-selected Extremely Red Objects'' (IERO; Yan et al. 2004). Using the two broad-band data taken by NIC3 in the HUDF (Thompson et al. 2005), YW04b identified three objects that are missing from the ACS images. Subsequent analysis using the Spitzer IRAC data suggest that two of them are red galaxies at $z\\approx 2$--3 without much on-going star formation (Yan et al. 2004; but also see Mobasher et al. 2005 and Chary et al. 2006), while the third one (ID No.3) is less conclusive because it is blended with other sources in the IRAC image and thus its photometry is uncertain. Bouwens et al. (2004) used the same data set, and pushed to a fainter limit and found a few additional candidates. Bouwens et al. (2005) further included all available deep NIC3 imaging data and extended their search to $z\\approx 10$, although no conclusive answer was obtained. Using these results, Bouwens et al. (2008; hereafter B08) derived constraints to the UV LF of galaxies at $z\\approx 7$--10, and argued that the LF evolves strongly and continues the dimming trend in $L^*$ (as proposed in B06) to higher redshifts. As a result, the number density of galaxies, and hence the UV luminosity density in the earlier universe should be significantly lower than at a later time. Meanwhile, the search for gravitationally lensed galaxies around foreground galaxy clusters have resulted in some remarkable success. Kneib et al. (2004) first discovered such an object that is likely at $z>6$. While no precise redshift was obtained, extensive optical and IR observations suggest that it is at $z\\approx 7$. Bradley et al. (2008) reported a very bright, highly probable candidate at $z\\approx 7.6$. Zheng et al. (2009) reported three new $z\\approx 7$ candidates from the same campaign. Richard et al. (2008) have found 12 candidates at similar redshifts. These results, however, still do not allow us put a strong constraint to the number density of galaxies at these redshifts, although they indeed prove that very high-redshift galaxies much fainter than our current detection limits (should there be no lensing magnification) do exist. There are also some other evidence from the search for \\Lya emitters around foreground clusters that supports similar conclusion (Stark et al. 2007). While the above results start to give us some meaningful constraints to the faint-end of the LF at $z\\gtrsim 7$, there is still only limited constraint to the bright-end where deep wide-field surveys are required. To date, most such surveys have null detection (e.g., Willis et al. 2008; Stanway et al. 2008; Sobral et al. 2009) or are uncertain (e.g. Hickey et al. 2009). One exception that has produced positive detections is the Subaru CCD survey for LAEs at $z=7.0\\pm0.1$, which has resulted in the redshift record of $z=6.96$ (Iye et al. 2006). However, even such LAEs can only constrain the faint population, as the vast majority of them are pure emission-line objects and are not seen in continuum. Another exception is that of Ouchi et al. (2009), which surveyed $\\sim 1500$~arcmin$^2$ in two fields and resulted in 22 $z\\approx 7$ candidates to $AB\\sim 26.0$~mag. Recently, Capak et al. (2009) reported three very bright candidates at $z\\approx 7$ at $J\\sim 23$~mag, however, their nature is still uncertain. Nearly all these studies (with the exception of Capak et al. 2009) in the very high-redshift frontier have claimed that their results are consistent with the strong declining evolution of the LF with respect to increasing redshift as suggested by B07 and B08. If this is indeed what the universe behaves, we are facing a dilemma: on one hand, the hydrogen reionization, which is now well constrained that must have started at $z\\approx 10$, requires a large amount of strong UV emitting sources, and the significant stellar mass density measured at $z\\approx 6$ also strongly suggests very active star formation activities at $z\\gtrsim 8$ and above; on the other hand, the limited number of observations to search for galaxies beyond $z\\approx 7$ indicates a strong declining number density of galaxies at higher redshifts. To reconcile these seemingly conflict results, a more decisive survey for galaxies at $z\\gtrsim 7$ is in demand. The Wide Field Camera 3 (WFC3) recently installed to \\hst\\, has provided a unique opportunity for the study of the universe at very high redshifts. The IR channel of this camera has a factor of $6.4\\times$ larger field-of-view (FOV) and an order of magnitude higher Q.E. as compared to NIC3, making it the most powerful tool in detecting galaxies at $z\\approx 7$ and beyond. For this reason, the first set of deep data that it took has already inspired four papers to appear at the arXiv preprint service within one week after the data were made public (Bouwens et al. 2009; Oesch et al. 2009; Bunker et al. 2009; McLure et al. 2009). All these new results, however, seem to reiterate that the number density of galaxies rapidly declines when we look back further in time. Here we present our results based on an independent reduction and analysis of these data. Our effort has resulted in more candidate galaxies at $z\\gtrsim 7$ than others, and, for the first time, a large sample of highly probable candidate galaxies at $z\\approx 10$. We will show that our analysis has led to a completely new, although still tentative conclusion about the formation and evolution of galaxies in the early universe. Our paper is organized as following. We briefly describe the WFC3 IR instrument and its observations of the HUDF in \\S 2, and give the details of our data reduction in \\S 3. Our photometry and catalog construction is described in \\S 4. The candidate selection and the dropout samples are presented in \\S 5. We discuss the implications of our results in \\S 6, followed by a summary in \\S 7. For simplicity, we denote the ACS passbands F435W, F606W, F775W, and F850LP as $B_{435}$, $V_{606}$, $i_{775}$, and $z_{850}$, respectively, and denote the three WFC3 IR passbands F105W, F125W and F160W as $Y_{105}$, $J_{125}$ and $H_{160}$, respectively. All magnitudes quoted are in AB system. Throughout the paper, we use the following cosmological parameters: $\\Omega_M=0.27$, $\\Omega_\\Lambda=0.73$ and $H_0=71$~km~s$^{-1}$~Mpc$^{-1}$. ", "conclusions": "" }, "0910/0910.3074_arXiv.txt": { "abstract": "We report the discovery of 31.18\\un{ms} pulsations from the \\integ~source \\igr~using the Rossi X-ray Timing Explorer (\\xte). This pulsar is most likely associated with the bright Chandra \\xray~point source lying at the center of \\snr, a previously unrecognised Galactic composite supernova remnant with a bright central non-thermal radio and \\xray~nebula, taken to be the pulsar wind nebula (PWN). \\psr~is amongst the most energetic rotation-powered pulsars in the Galaxy, with a spin-down luminosity of $\\dot E = 5.1 \\times 10^{37}$ erg~s$^{-1}$. In the rotating dipole model, the surface dipole magnetic field strength is $B_s = 1.1 \\times10^{12}$~G and the characteristic age $\\tau_c \\equiv P/2\\dot P = 12.7$~kyr. The high spin-down power is consistent with the hard spectral indices of the pulsar and the nebula of $1.22\\pm0.15$ and $1.83\\pm0.08$, respectively, and a 2--10\\un{keV} flux ratio $F_{PWN}/F_{PSR} \\sim 8$. Follow-up Parkes observations resulted in the detection of radio emission at 10 and 20\\un{cm} from \\psr~at a dispersion measure of $\\sim$ 560 cm$^{-3}$ pc, which implies a relatively large distance of 10 $\\pm$ 3\\un{kpc}. However, the resulting location off the Galactic Plane of $\\sim$ 280\\un{pc} would be much larger than the typical thickness of the molecular disk, and we argue that \\snr~lies at a distance of $\\sim$ 7\\un{kpc}. There is no gamma-ray counterpart to the nebula or pulsar in the Fermi data published so far. A multi-wavelength study of this new composite supernova remnant, from radio to very-high energy gamma-rays, suggests a young ($\\lesssim$ 10$^{3}$ yr) system, formed by a sub-energetic ($\\lesssim 10^{50}$ ergs), low ejecta mass (M$_{\\rm ej} \\sim 3$ \\msun) SN explosion that occurred in a low-density environment ($n_0 \\sim$ 0.01 cm$^{-3}$). ", "introduction": "\\label{s:intro} The number of known supernova remnants (SNRs) in the Galaxy has increased significantly over the last few years, mainly due to a new generation of radio and \\xray~instruments of unprecedented sensitivities and angular resolutions. In particular, these observations, combined with Galactic Plane surveys \\cite[\\eg][]{c:brogan06}, and targeted observations \\cite[\\eg][]{c:gelfand07,c:gaensler08} have increased the fraction of SNRs found to harbor an energetic pulsar powering a wind nebula (the so-called composite SNRs). Moreover, deep observations toward compact PWNe have proven to be successful in detecting their powering pulsars \\citep{c:camilo09,c:gotthelf09}, providing constraints on their energetics, spin evolution, and birth parameters. Along with the current generation of high/very-high energy (VHE; $>$ 100\\un{GeV}) instruments, these observations are of prime importance for understanding the structure and evolution of these sources, and the underlying acceleration mechanisms which occur close to the pulsar, and at the relativistic and non-relativistic shock fronts bounding the PWN and the host SNR \\citep{c:gs06}. The soft $\\gamma$-ray source \\igr~was discovered in a deep \\integ/\\ibisg~mosaic of the Circinus region as a persistent source at the mCrab level \\citep{c:keek06}. A \\swi/XRT survey of \\integ~sources located \\igr~to 4\\s~in 2--10\\un{keV} \\xrays, but no conclusion was reached on its nature \\citep{c:malizia07}. A follow-up \\chandra~survey of unidentified IGR sources reported a PWN within a $\\sim$ 3\\m~diameter nearly circular emission nebula \\citep{c:tomsick09}. These authors presented the 0.3--10\\un{keV} spectrum of the total emission by an absorbed power-law with a relatively hard photon index ($\\Gamma$ = 1.82 $\\pm$ 0.13) and a large column density \\nh~$\\sim$ 3 $\\times$ 10$^{22}$ cm$^{-2}$. In this paper, we report the discovery with the Rossi X-ray Timing Explorer (\\xte) of 31.18\\un{ms} pulsations toward \\igr, and the radio detection in follow-up observations using the Parkes telescope. We present a multi-wavelength study of this new Galactic composite SNR, \\snr, using radio, X-ray, and gamma-ray data. \\psr~is one of the most energetic in the Galaxy and powers a wind nebula whose broadband non-thermal synchrotron emission is measured in radio and \\xrays. We present a spatial and spectral analysis that shows evidence for a young SNR, formed from a sub-energetic, low ejecta mass SN explosion that occurred in a low-density environment. We also discuss the implications of the lack of high-energy (HE) gamma-ray emission in the \\fermi~data. ", "conclusions": "Even though some assumptions have been made in the previous estimates (\\eg~constant PWN magnetic field and PSR spin-down power), they give valuable insight on the nature of this new Galactic composite SNR. \\snr~harbors a highly energetic 31.18\\un{ms} pulsar, \\psr, which powers a wind nebula, both lying at the center of the host shell. All of the existing multi-wavelength observations suggest it is a young SNR ($\\lesssim$ 10$^{3}$ yr), and most likely distant ($>$ 5\\un{kpc}). However, many questions still remain to be answered. First, the distance is not very well constrained, though a large distance is favored by the dispersion measure of the radio pulse emission and by the large \\nh~measured with \\chandra. High-resolution observations of the ISM tracers such as HI and $^{12}$CO are then warranted to assess the surrounding medium properties. Moreover, the break measured at 6\\un{keV} and the spectral softening at increasing distances in the PWN need to be confirmed, and the SNR \\xray~spectrum needs to be investigated with more \\xray~data. Nevertheless, \\snr~falls into the emerging class of ``multi-wavelength'', young ($\\tau \\lesssim$ a few 10$^{3}$ yr) and composite SNRs, harboring very energetic PSRs and wind nebulae shining in radio, \\xrays~and potentially in HE/VHE \\gammarays. The list includes Kes~75 and G21.5$-$0.9 \\citep{c:gotthelf00,c:camilo06,c:bietenholz08,c:djannati08}, and more recently G0.9+0.1 \\citep{c:aharonian05a} and HESS~J1813$-$178 \\citep{c:aharonian05b}, whose long-expected PSRs have recently been discovered \\citep{c:camilo09,c:gotthelf09}. It is of interest to note that \\fermi/LAT has not detected a bright source \\citep{c:fermi_soucat}, nor pulsar \\citep{c:fermi_psr1,c:fermi_psr2} coincident with \\snr, although, at first glance, energetic PSRs should be the most easily detectable sources. However, not all of the above-mentioned young and energetic PSRs have been detected by \\fermi. From the first \\fermi~catalog of gamma-ray PSRs, the sensitivity for a blind search of pulsed emission in the Galactic Plane is conservatively taken to be 2 $\\times$ 10$^{-7}$ cm$^{-2}$ s$^{-1}$ above 100\\un{MeV} \\citep{c:fermi_psr2}. Assuming a spectrum similar to that of PSR~J1833$-$1034 associated with G21.5$-$0.9, \\psr~features a maximal efficiency $\\eta =\\,L_{\\gamma}$/\\edot~of $\\sim$ 1 $d_{7}^2$ \\%. This is close to what is measured from other young PSRs (see Fig. 6 of \\citet{c:fermi_psr2}), and argues in favor of a fairly large distance to the source, as outlined in section \\ref{s:discu}. Further HE/VHE observations of \\snr~will certainly provide important constraints both on the PSR gamma-ray spectrum and on the PWN magnetic field strength." }, "0910/0910.1592_arXiv.txt": { "abstract": "Here we show that the overabundance of ultra-luminous, compact X-ray sources (ULXs) associated with moderately young clusters in interacting galaxies such as the Antennae and Cartwheel can be given an alternative explanation that does not involve the presence of intermediate mass black holes (IMBHs). We argue that gas density within these systems is enhanced by the collective potential of the cluster prior to being accreted onto the individual cluster members and, as a result, the aggregate X-ray luminosity arising from the neutron star cluster members can exceed $>10^{39}\\;{\\rm erg s^{-1}}$. Various observational tests to distinguish between IMBHs and accreting neutron star cusps are discussed. ", "introduction": "Over the years, the existence of two distinct populations of black holes has been established beyond a reasonable doubt. Supermassive black holes, $M > 10^6 \\, M_\\sun$, are inferred in many galactic centers \\citep{1995ARAA..33..581K,1998AJ....115.2285M}, while stellar mass black holes, $M \\sim 1-10 \\, M_\\sun$, have been identified by their interaction with companion stars \\citep{2006csxs.book..157M}. The situation at intermediate masses, $M \\sim 10^2 - 10^5 M_\\sun$, is still uncertain despite recent evidence for mass concentrations within the central regions of some globular clusters \\citep{2005ApJ...634.1093G,2007ApJ...661L.151U,2008ApJ...676.1008N}. This evidence remains controversial, partly because the velocity dispersion profiles can be reproduced without invoking the presence of an intermediate mass black hole (IMBH) \\citep{baum1, baum2,2009arXiv0905.0627A}. Recently, some evidence has arisen for the presence of IMBHs in moderately young star clusters, where ultra-luminous, compact X-ray sources (ULXs) have been preferentially found to occur \\citep{2001ApJ...554.1035F,2008AIPC.1010..357T}. Their high luminosities have been interpreted as imprints of IMBHs \\citep{2004Natur.428..724P}, rather than binaries containing a normal stellar mass black hole \\citep{2006ApJS..166..211Z}. In this {\\it Letter}, we present an alternative explanation for the overabundance of ULXs associated with young clusters in galaxies such as the Antennae and Cartwheel. In this new paradigm, the accretion of gas by the collective star cluster potential moving through the merging medium is strongly enhanced relative to the individual rates and, as a result, the aggregate X-ray luminosity arising from the neutron star cluster members can exceed $>10^{39}\\;{\\rm erg s^{-1}}$. Much of the effort herein will be dedicated to understanding the conditions by which the collective potential of a star cluster is able to accrete gas with highly enhanced rates and its effect on the integrated accretion luminosity of the neutron star cluster members. The conditions found in systems such as the Antennae galaxy, as we will argue, are favorable for this type of mechanism to operate effectively and produce an overabundance of ULXs. ", "conclusions": "Many studies have been focused on the flow around compact stars with a point mass potential. Although clusters have much larger masses than individual stars, their potential is relatively shallow. In this paper we consider the efficiency of accretion in these cluster potentials, and show that when the sound speed or the relative velocity of the ambient medium is less than the central velocity dispersion of the cluster, the collective potential alters the local gas flow before the gas is accreted onto the individual stars within the cluster. Accretion onto these dense stellar cores at the inferred rate can lead to the onset of ULX sources. While there are no stellar clusters observed in the Galactic disk which bear these anticipated properties (the relative velocity of the halo clusters to the interstellar medium is in the range of 100 km/s), observations of several cluster knots in the Antennae indicate intracluster relative velocities that comparable to the central velocity dispersions \\citep{2005AJ....130.2104W}. Based on the results of the current work, we show that accretion by individual compact stars in the centers of such systems is enhanced greatly relative to their rate of accretion directly from the ambient gas, and conclude that this process may be relevant for explaining the origin of ULX sources in these extraordinary clusters. Illumination of nearby gas clouds by these sources may also lead to reprocessed infrared, optical and ultraviolet emission. Finally, the sources may leave trails of denser and likely hotter gas behind them as they plough through the gas. A way to distinguish between an IMBH \\citep{2004Natur.428..724P} and an accreting neutron star cusp is via time-dependent observations. The emission from a relativistic region of a IMBH might vary on time-scales of seconds. The emission of a large number of statistically independent black holes and neutron stars should be considerably less variable: $\\Delta t \\leq r_{\\rm c}/c_{\\rm s}\\sim 10^4$ yr. Observations find that a handful of the X-ray sources in the Antennae galaxy are indeed variable albeit on timescales that are larger than a few years \\citep{2006ApJS..166..211Z}. Such variability might be explained if only a moderate fraction of the compact stars dominate the total luminosity. Such compact isolated accretors will probably have unusual time-variability properties as their discs may be much larger than the typical discs of X-ray binaries, and indeed they are missing the perturbing influence of the secondary. On the other hand, accretion disc feeding in these sources will be variable itself, leading to variability on variety of time-scales. Although accretion disk spectra are notoriously difficult to calculate from first principles, an IMBH and a cluster core may also have observably different spectra. It has been suggested that a cool multicolor disk spectral component, might indicate the presence of an IMBH \\citep{2003ATel..212....1M}. This is understood as following. The larger mass of the BH accretor, the lower the temperature of the inner edge of the disk, which scales as $T_{\\rm i} \\propto (M_{\\rm BH}\\dot{M}/r_{\\rm i})^{1/4} \\propto \\dot{M}^{1/4}M_{\\rm BH}^{-1/2}$ for a simple thin-disk model. Since for our model, individual neutron star sources have $T_{\\rm i}\\sim 1$ keV (consistent with observations), then an IMBH might have $T_i \\sim 30$ eV. Thus, an association of the ULXs with an IMBH, as opposed to a accreting distribution of compact remnants, could be made on the basis of a very soft observed spectral component. We know of no reported observations of such components." }, "0910/0910.1906_arXiv.txt": { "abstract": "Dark energy must cluster in order to be consistent with the equivalence principle. The background evolution can be effectively modeled by either a scalar field or by a barotropic fluid. The fluid model can be used to emulate perturbations in a scalar field model of dark energy, though this model breaks down at large scales. In this paper we study evolution of dark energy perturbations in canonical scalar field models: the classes of thawing and freezing models. The dark energy equation of state evolves differently in these classes. In freezing models, the equation of state deviates from that of a cosmological constant at early times. For thawing models, the dark energy equation of state remains near that of the cosmological constant at early times and begins to deviate from it only at late times. Since the dark energy equation of state evolves differently in these classes, the dark energy perturbations too evolve differently. In freezing models, since the equation of state deviates from that of a cosmological constant at early times, there is a significant difference in evolution of matter perturbations from those in the cosmological constant model. In comparison, matter perturbations in thawing models differ from the cosmological constant only at late times. This difference provides an additional handle to distinguish between these classes of models and this difference should manifest itself in the ISW effect. ", "introduction": "Various observation have confirmed that the expansion rate of the universe is accelerating \\cite{obs_proof}. These observations include those of Supernova type Ia \\cite{nova_data}, observations of Cosmic Microwave Background \\cite{boomerang,wmap_params} and large scale structure \\cite{sdss}. The accelerated expansion of the universe can be explained by introducing a cosmological constant $\\Lambda$ in the Einstein's equation \\cite{ccprob_wein,review3}. However, the cosmological constant model is plagued by the fine tuning problem \\cite{ccprob_wein}. This has motivated the study of dark energy models to explain the current accelerated expansion of the universe (for reviews see \\cite{DEreview}). An alternative to the cosmological constant model is to assume that this accelerated expansion is driven by a canonical scalar field with a potential $V(\\phi)$, namely the quintessence field \\cite{quint1,expo,linear,quadratic,invphi,invexpo}. There exists another class of string theory inspired scalar field dark energy models known as tachyon models \\cite{tachyon1,2003PhRvD..67f3504B} and there are models which allow $w<-1$ are known as phantom models \\cite{2002PhLB..545...23C}. Phantom type dark energy can also be realized in a scalar tensor theory of gravitation \\cite{STG}. Other scalar field models include k-essence field \\cite{2001PhRvD..63j3510A}, branes \\cite{brane1} and fluid models like the Chaplygin gas model and its generalizations \\cite{chaply}. There are also some phenomenological models \\cite{water}, field theoretical and renormalization group based models (see e.g. \\cite{tp173}), models that unify dark matter and dark energy \\cite{unified_dedm1}, holographic dark energy models \\cite{HGDE}, QCD dark energy \\cite{qcddark} and many others like those based on horizon thermodynamics (e.g. see \\cite{2005astro.ph..5133S}). Different models of dark energy which have the same background evolution are indistinguishable purely from distance measurements. Evolution of perturbations in these models is expected to break this degeneracy. The Integrated Sachs Wolfe (ISW) effect can distinguish a cosmological constant from other models of dark energy, especially ones with a dynamical dark energy \\cite{ddw}. Dark energy perturbations have been extensively studied in the linear approximation \\cite{weller_lewis,bean_dore,depert,gordonhu}. Perturbations in dark energy affect the low $l$ quadrupole in the CMB angular power spectrum through the ISW effect \\cite{weller_lewis,bean_dore}. It was shown in Ref.\\cite{weller_lewis} that dark energy perturbations affect the low $l$ quadrupole in the CMB angular power spectrum through the ISW effect. For models with $w>-1$ this effect is enhanced while for phantom like models it is suppressed. In these models dark matter perturbations and dark energy perturbations are anti-correlated for large effective sound speeds. This anti-correlation is a gauge dependent effect \\cite{bean_dore}. There are several other studies of perturbations in dark energy \\cite{chpgas_pert}, including some that deal with evolution of spherical perturbations \\cite{sph_coll,mota}. For canonical scalar field dark energy, the perturbations in matter are enhanced by the presence of dark energy perturbations in comparison with smooth dark energy model \\cite{ujs}. The matter perturbations in fluid models are suppressed compared to corresponding homogeneous dark energy scenario \\cite{hkj}. As long as the speed of propagation of perturbations `$c_s^2$' is positive the evolution of matter perturbations is indistinguishable from a smooth dark energy model. This is true for scales smaller then the Hubble radius. Dark energy perturbations in a fluid model with an appropriate $c_s^2$ emulate that of a scalar field model very well below the Hubble scale but start to differ at larger scales. Therefore the fluid model is not a good approximation at these scales. This also implies that the growth of perturbations at large scales depends on the details of the model even though the background evolution is the same. A separate analysis is therefore required for every model. In this paper we consider different scalar field models to study evolution of dark energy perturbations. We consider two different types of potentials, classified as '{\\it thawing}' and '{\\it freezing}' in Ref. \\cite{limitsofq}. For potentials with a thawing behavior, the scalar field is frozen at early times and starts to roll down the potential at late time. Hence the equation of state of dark energy starts near $w=-1$ at early times to $w>-1$ at late times. In contrast, in the case of potentials with freezing behavior, the scalar field rolls down the potential and approaches the minimum of the potential, with the equation of state going from $w>-1$ to freezing at $w=-1$ at late times. Due to the different evolution of the equation of state, it is expected that the perturbations in dark energy evolve differently in these classes. For freezing potentials, at early times, the equation of state deviates from $w=-1$ hence the perturbations in dark energy are expected to be enhanced. Whereas, in thawing potential scenarios, dark energy perturbations become significant only at late times. Since dark energy perturbations enhance perturbations in nonrelativistic matter, models with freezing type potentials will have more enhanced perturbations in matter than those with thawing type behavior. This can be an additional tool (apart from distance measurements) to distinguish between these types of models. The paper is organized as follows. In Section \\ref{sec::fg} we discuss evolution of perturbations in scalar field potentials in the thawing and freezing classes. The results are summarized in the concluding Section \\ref{sec::concl}. ", "conclusions": "\\label{sec::concl} In this paper, we analyze the growth of perturbations in scalar field dark energy scenarios. The assumption that the distribution of dark energy (with $w \\neq -1$) is homogeneous at all length scales is inconsistent with the equivalence principle. On length scales comparable to or greater than the Hubble radius, the perturbations in dark energy can become comparable to perturbation in matter if $w_{de} \\neq -1$. For scales smaller than the Hubble radius, perturbations in dark energy can be neglected in comparison with the perturbation in matter at least in the linear regime. Hence any deviations from the cosmological constant model can be explored by assuming dark energy as a homogeneous component and the scalar field models can be approximated well by parameterized fluid models. However, on much larger scales, if the equation of state parameter deviates from -1, then perturbations in dark energy do influence matter power spectrum. For canonical scalar fields, a clustering dark energy enhances matter perturbations as compared to the corresponding homogeneous dark energy scenario. This enhancement more pronounced at scales larger than the Hubble radius. In particular, we study two broad classes of canonical scalar field models, namely the thawing and freezing models. The equation of state of dark energy evolves differently in these two classes of models. This affects the growth of perturbations in these different types. For fast rolling, freezing type models, the dark energy equation of state deviates away from that of a cosmological constant at early times and freezes to $w=-1$ at late times. In slow roll thawing models, the scalar field remains at a constant $w=-1$ and starts to deviate away from this value at late times. The present day observations, which are based on distance measurements, cannot distinguish between these models. We studied evolution of matter perturbations in the presence of a clustering dark energy. The models we have considered in this paper are within the range allowed by distance measurement observations. For corresponding homogeneous dark energy models and clustering dark energy models, there are significant changes in the matter density contrast evolution. All the canonical scalar field models studied here show an enhancement in matter perturbations if dark energy is perturbed. Although we have not considered phantom like models in this paper, it is worth mentioning that in these models, dark energy perturbations suppress matter perturbations. In general, freezing type models have a higher rate of growth of density contrast at early times. This is due to the fact the in these models, the equation of state of dark energy is further away from that of a cosmological constant. Hence dark energy perturbations play a prominent role at earlier times. For thawing type models, in general, dark energy perturbations affect the matter perturbations at late times. For most of the evolution, the matter perturbations remain close to those in cosmological constant model and being to deviate as the field begins to thaw. There are significant deviations in the way density contrast grows, not only between various models but also from the concordant cosmological constant model. Apart from different way matter perturbations grow in the freezing and thawing classes of models, models within the same class also have variation in the behavior of the density contrast. Observable changes in the angular power spectrum at large scales are limited by the cosmic variance and therefore CMB data alone will be insufficient to distinguish between these models. Since the scales at which dark energy perturbations are relevant are large, the primary contribution to CMB anisotropies is through the ISW effect. Therefore it is important to study the ISW cross correlation with large scale structure indicators." }, "0910/0910.4238_arXiv.txt": { "abstract": " ", "introduction": "The role of the gas in the formation and evolution process of early-type galaxies is still not fully understood. For example, recent studies have shown a large complexity in the gas structures in these systems (e.g. Morganti et al. 2006, Sarzi et al. 2006, Combes et al 2007) despite their often unspectacular optical appearance. In addition, some sources show nuclear activity while others do not. Cold-gas structures represent a fossil record of the formation and evolution of early-type galaxies. In particular, gas found on kiloparsec scales can be used to trace the evolution of the host galaxy (e.g. major merger vs. small accretions). In that respect, neutral hydrogen is an important tracer of these events as it often extends the dust and optical (disk) structure by a factor of two or more. Close to the centre of galaxies ($<100$~pc), (cold) gas also plays a crucial role as it can provide the fuel that is needed to make the central black hole active. It is often believed that mergers are important in driving gas to the centre (see e.g. Hibbard \\& van Gorkom 1996, Barnes 2002), but recent studies of radio galaxies have shown that the activity in some galaxies may be associated with the (slow) accretion of gas from the ISM/IGM (e.g. Best et al. 2005). Despite the need for large statistical investigations to understand (and classify) the different mechanisms at work (e.g. mergers, interactions, accretion, AGN activity etc.), it is indispensable to study close-by objects in great detail with very high linear resolution. For example, the circumnuclear region around the black hole ($< 100$~pc) can only be resolved in the most nearby galaxies to a degree that is needed to understand the accretion/fueling process. By far the closest radio-loud early-type galaxy is Centaurus~A (NGC~5128) at a distance of only 3.8~Mpc\\footnote{At this distance 1\\arcmin ~corresponds to $\\sim 1.1$~kpc.} (Harris 2009). Cen~A has been studied in all possible wavelength regimes with very high linear resolution (for an overview see the review by Israel (1998) and the other contributions to this volume). However, it is crucial to compare Cen~A with other sources to check whether it is a typical example of its class, or whether it is unusual w.r.t. other radio-loud, low-luminosity sources. In this paper we discuss whether Cen~A is special as seen from the neutral hydrogen perspective. That is, do the Cen~A properties (such as e.g. \\HI mass, morphology, kinematics etc.) differ from other early-type galaxies and does Cen~A share properties with radio galaxies that have comparable luminosity? In Sect.~2 we give a brief description of the \\HI morphology and kinematics on kpc and sub-kpc scales. Section~3 compares Cen~A with other nearby early-type galaxies and in Sect.~4 Cen~A is compared to a complete sample of nearby radio galaxies. We summarize our comparison in Sect.~5. ", "conclusions": "In order to understand Centaurus~A in the context of galaxy formation and evolution, we have compared the \\HI properties of Cen~A with early-type and radio galaxies. The \\HI mass, its distribution and the mainly settled kinematics is commonly found in other early-type/radio galaxies. The current phase of AGN activity is not connected to the recent merger which is in line with recent results for a sample of radio galaxies. The absorption against the nucleus is red- and blueshifted with respect to the systemic velocity and is in agreement --- as is also the case in other sources --- with a circumnuclear \\HI disk/torus structure. Hence, Centaurus~A seems to be --- from an \\HI perspective --- a typical galaxy of its class." }, "0910/0910.2714_arXiv.txt": { "abstract": "{Current stellar population models have arguably the largest uncertainties in the near-IR wavelength range, partly due to a lack of large and well calibrated empirical spectral libraries. In this paper we present a project, which aim it is to provide the first library of luminosity weighted integrated near-IR spectra of globular clusters to be used to test the current stellar population models and serve as calibrators for the future ones. Our pilot study presents spatially integrated $K$-band spectra of three old ($\\ge$10~Gyr) and metal poor ([Fe/H]$\\sim$~--1.4), and three intermediate age (1~--~2~Gyr) and more metal rich ([Fe/H]$\\sim$~--0.4) globular clusters in the LMC. We measured the line strengths of the \\na\\/, \\ca\\/ and \\co\\/ absorption features. The \\na\\/ index decreases with the increasing age and decreasing metallicity of the clusters. The \\dco\\/ index, used to measure the \\co\\/ line strength, is significantly reduced by the presence of carbon-rich TP-AGB stars in the globular clusters with age $\\sim$1~Gyr. This is in contradiction with the predictions of the stellar population models of Maraston (2005). We find that this disagreement is due to the different CO absorption strength of carbon-rich Milky Way TP-AGB stars used in the models and the LMC carbon stars in our sample. For globular clusters with age $\\geq$2~Gyr we find \\dco\\/ index measurements consistent with the model predictions.} ", "introduction": "\\label{sec:motivation} Since the 90's, the interpretation of the integrated light of galaxies (in the nearby universe or at high redshift) heavily relies on evolutionary population synthesis (EPS) models. Such models were pioneered by \\citet{tinsley80} and the method has been seriously extended since then \\citep[e.g.][]{bc93,worthey94,vazdekis96,fioc97,starburst99, maraston05, schiavon07}. They are used to determine ages, element abundances, stellar masses, stellar mass functions, etc., of those stellar populations that are not resolvable into single stars with today's instrumentation, i.e. most of the universe outside the Local Group. To build such EPS models we use simple stellar populations (SSP). There are two essential advantages of focusing on SSPs. First, SSPs can be reliably calibrated. They can be compared directly with nearby globular cluster (GC) data for which accurate ages and element abundances are independently known from studies of the resolved stars. This step is crucial to fix the stellar population model parameters that are used to describe model input physics and which cannot be derived from first principles (e.g., convection, mass loss and mixing). Second, SSPs can be used to build more complex stellar systems. Systems made up by various stellar generations can be modelled by convolving SSPs with the adopted star formation history \\citep[e.g.][] {kodama97, bc03}. Models describing accurately the integrated light properties, including medium to high resolution spectra and/or line-strength indices, are and will be our main tool to investigate and analyse the star-formation history over cosmological time-scales. This approach has worked well in the optical spectroscopic regime and has led to well calibrated models \\citep[e.g.][]{tmb03,bc03,maraston05}. With the application of such models to observed spectra we derive reasonable estimates of the main stellar population parameters (age, chemical composition and M/L ratio) in the nearby universe \\citep[e.g.,][]{kun00,trager00,thomas05,cappellari06,sb07} as well as at higher redshifts \\citep[e.g.,][]{bernardi05,maraston05,sb09}. Of course, uncertainties remain due to the degeneracy of age and metallicity effects in the optical wavelength range \\citep[e.g.,][] {worthey94}. The integrated near-IR light in stellar populations with ages $\\geq$ 1 Gyr is dominated by one stellar component, cool giant stars, whose colour and line indices are mainly driven by one parameter: metallicity \\citep{frogel78}. Near-IR colours and indices also have the advantage of being more nearly mass-weighted, i.e. the near-IR mass-to-light ratio is closer to one \\citep [see e.g.,][]{worthey94}. So, by combining the optical as well as near-IR information one can resolve the currently remaining degeneracies between age and chemical composition, present in the models, and hope to gain a better understanding of star-formation histories. However, currently available stellar population models have arguably the largest uncertainties in the near-IR and thus it is of paramount importance to provide high-quality observational data to validate and improve the state-of-the-art models. Globular clusters in the Local Group are an ideal laboratory for this project since ample information from studies of the resolved stars is available. Yet, integrated spectroscopic observations of the Galactic GCs in the near-IR are very challenging due to their large apparent sizes on the sky. The Large Magellanic Cloud (LMC) and its globular cluster system, located about 50 kpc away, is a much better observational choice. It shows evidence for a very complex and still ongoing star formation activity. The LMC GCs have an advantage (for the scope of this project) with respect to Galactic GCs - they span a larger range in ages. Studies of the LMC globular cluster system show one old component with age $>$10~Gyr. After this time there was a ''dark age'' with just one cluster formed before a new burst of cluster formation that has started around $3-4$~Gyr ago \\citep{dacosta91}. A disadvantage is their lower metallicity. \\begin{center} \\begin{table*}[tdp] \\centering \\caption{\\label{tab:lmc_observations}Target globular clusters in the LMC -- observing log.} \\begin{tabular}{c c c c c c c } \\hline \\hline Name & UT date & $V$ & $(B-V)$ & SWB & Age & [Fe/H]\\\\ (1) & (2) & (3) & (4) & (5) & (6) & (7) \\\\ \\hline NGC\\,1754 & 2006 Nov 18 & 11.57 & 0.75 & VII & 10 & -1.42$^{a}$, -1.54$^{b}$\\\\ NGC\\,2005 & 2006 Nov 29 & 11.57 & 0.73 & VII & 10 & -1.35$^{a}$, -1.92$^{b}$, -1.80$^{c}$, -1.33$^{d}$\\\\ NGC\\,2019 & 2006 Dec 10 & 10.86 & 0.76 & VII & 10 & -1.23$^{a}$, -1.18$^{b}$, -1.37$^{c}$, -1.10$^{d}$\\\\ NGC\\,1806 & 2006 Dec 06 & 11.10 & 0.73 & V & 1.1& -0.23$^{b}$, -0.71$^{e}$ \\\\ NGC\\,2162 & 2006 Dec 05 & 12.70 & 0.68 & V & 1.1 & -0.23$^{b}$, -0.46$^{f}$\\\\ NGC\\,2173 & 2006 Dec 06 & 11.88 & 0.82 & V-VI & 2& -0.24$^{b}$, -0.42$^{f}$, -0.51$^{g}$\\\\ \\hline \\hline \\end{tabular} \\smallskip \\flushleft Notes: (1) Cluster name, (2) date of observation, (3) integrated $V$-band magnitude, (4) $(B-V)$ colour and (5) SWB type taken from \\citep{bica96,bica99}, (6) Age of the cluster in Gyr, based on the SWB type \\citep{frogel90}, (7) [Fe/H] derived using different methods: $^{a}$ \\citet{olsen98} -- slope of the RGB; $^{b}$ \\citet{olszewski91} -- low-resolution Ca\\,II triplet; $^{c}$ \\citet{johnson06} -- high-resolution Fe\\,I; $^{d}$~\\citet{johnson06} -- high-resolution Fe\\,II; $^{e}$ \\citet{dirsch00} -- Str\\\"{o}mgren photometry; $^{f}$ \\citet{groch06} -- low-resolution Ca\\,II triplet; $^{g}$~\\citet{muc08} -- high-resolution spectroscopy. \\end{table*} \\end{center} The goal of this project is to provide an empirical near-IR library of spectra for integrated stellar populations with ages $\\geq$ 1 Gyr, which will be used to verify the predictions of current SSP models in the near-IR wavelength range. Here we present the results from a pilot study of $K$-band spectra of 6 globular clusters in the LMC. The analysis of their $J$ and $H$-band spectra will be discussed in a separate paper. This paper is organised as follows: in Sect.~\\ref{sec:sample} we give details about the sample selection and the observing strategy. Sect.~\\ref{sec:obs} is devoted to the observations and data reduction. In Sect.~\\ref{sec:in_or_out} we discuss the cluster membership of the stars in our sample, in Sect.~\\ref{sec:indices} we describe the near-IR index measurement procedures. In Sect.~\\ref{sec:data_models_lmc} we make a comparison between the currently available stellar population models in the near-IR with our data, discuss the observed disagreements, and give potential explanations. Finally, in Sect.~\\ref{sec:conclusions} we give our concluding remarks. ", "conclusions": "\\label{sec:conclusions} The goal of this project is to provide an empirical spectral library in the near-IR for integrated stellar populations with ages $>$~1~Gyr, which will be used to test the current and calibrate the future stellar population models. In this paper we have presented the first results from a pilot study of the $K$-band spectroscopic properties of a sample of six globular clusters in the LMC. To validate the observational strategy, data reduction and analysis methods, we have selected from the catalogue of \\citet{bica99} three out of 38 GCs with SWB type VII to represent the old ($>$10~Gyr) and metal poor ([Fe/H]$\\sim$--1.4) population of the LMC, and three out of 71 clusters with SWB types V and VI to explore the properties of the intermediate age (1\\,--\\,3~Gyr) and more metal rich ([Fe/H]$\\sim$--0.4) component of the population. For each cluster our integrated spectroscopy covers the central $24\\arcsec\\times24\\arcsec$ and in most of the cases we have sampled about half the light. However, in order to better sample bright AGB stars, which are the most important contributors to the integrated cluster light in the near-IR, we have observed up to 9 of the brightest stars outside the central mosaics, but still within the tidal radii of the clusters, that have near-IR colours and magnitudes consistent with bright red giants in the observed clusters. We obtained integrated luminosity weighted spectra for the six clusters, measured the line strengths of \\na\\/, \\ca\\/ and \\co\\/ absorption features in the $K$-band and compared the strength of \\co\\/ with the stellar population models of \\citet{maraston05}. The observing strategy to cover at least the central half-light radius with a number of SINFONI pointings showed to be an efficient way in sampling the near-IR light of old ($>$\\,10\\,Gyr) clusters. For the intermediate age and sparse clusters, which are dominated by just a few very bright stars, observing a central mosaic plus a number of the brightest stars in the vicinity of the cluster is a better choice for optimal cluster light sampling. The availability of high spectral resolution spectroscopy greatly helps the differentiation of the cluster member stars from the LMC field population. In intermediate age clusters the largest amount of light originates from oxygen (M-type) and carbon-rich (C-type) AGB stars. Different ratios of the contributions by these two types of stars can lead to significant changes in the near-IR \\co\\/ line strength. According to our observations, when the C-type stars contribution peaks (at $\\sim$1\\,Gyr) the observed CO line strength is weak and then increases fast to reach its maximum for clusters with age $\\sim$2\\,Gyr. It is important to note that a weak line strength of \\co\\/ does not mean that there is less CO in these clusters/stars. The indices, which are used to describe the line strengths, take also into account the continuum shape, which in the case of carbon-rich stars is severely affected by typical for this type of stars absorption features and thus the resultant index value is low. The comparison of our data with the stellar population models of \\citet{maraston05}, in terms of CO line strength, shows a disagreement for the youngest clusters in our sample. For clusters with age $\\sim$1\\,Gyr the models predict maximal CO line strength, while we observe the opposite: the CO strength is significantly weaker. At the same time, literature data of the integrated colours of the clusters are consistent with these models. We explain these discrepancies as due to the different origin of the C-type stars used to calibrate the models and the ones in our data sample. The stars, used for model calibration, are Milky Way carbon stars, while our carbon stars are born in the LMC. We support this scenario with Fig.~\\ref{fig:co_jk_cstars}, where we show that carbon rich stars in the Milky Way and LMC, which have similar $(J-K)$ colour, have very different CO line strengths. This deserves further investigation and hopefully the next generation of carbon star models \\citep[e.g.][]{aringer09} will help us to elucidate whether a systematic effect of the metallicity on CO indices is expected, or the found discrepancy is due to the small sample of individual observations of variable stars in existing libraries. The near-IR \\na\\/ index shows a dependance on the age of the clusters -- it is decreasing with an increasing age. The combination between optical and near-IR spectral indices seems to offer possibilities to break the age-metallicity degeneracy, but more accurate and detailed stellar population models are necessary in the near-IR wavelength range. These models are of paramount importance, when studying the spatially resolved stellar populations of nearby galaxies. Adaptive optics assisted observations allow for the best correction of the Earth's atmosphere perturbing effects when observing in the near-IR. First attempts to explore galaxy evolution via spatially resolved near-IR spectroscopy \\citep[e.g.][]{az08,davidge08,nowak08} have shown the need for a better understanding of the properties of stellar populations in this wavelength range. The availability of detailed and reliable stellar population models in the near-IR will open up a new window to the exploration of galaxy formation and evolution. The final integrated spectra of the six LMC globular clusters will be made available via the Strasbourg astronomical Data Center (CDS)." }, "0910/0910.0711_arXiv.txt": { "abstract": "We have used a statistical technique ``Shuffle\" \\citep {bhav, bharad2} in seven nearly two dimensional strips from the Sloan Digital Sky Survey Data Release Six (SDSS DR6) to test if the statistically significant length scale of filaments depends on luminosity, colour and morphology of galaxies. We find that although the average filamentarity depends on these galaxy properties, the statistically significant length scale of filaments does not depend on them. We compare it's measured values in SDSS against the predictions of $\\Lambda$CDM N-body simulations and find that $\\Lambda$CDM model is consistent with observations. The average filamentarity is known to be very sensitive to the bias parameter. Using $\\Lambda$CDM N-body simulations we simulate mock galaxy distributions for SDSS NGP equatorial strip for different biases and test if the statistically significant length scale of filaments depends on bias. We find that statistically significant length scale of filaments is nearly independent of bias. This result is possibly related to the fact that statistically significant length scale of filaments is nearly the same for different class of galaxies which are differently biased with respect to underlying dark matter distribution. The average filamentarity is also known to be dependent on the galaxy number density and size of the samples. We use $\\Lambda$CDM dark matter N-body simulations to test if the statistically significant length scale of filaments depends on number density of galaxies and size of the samples. Our analysis shows that the statistically significant length scale of filaments very weakly depends on these factors. Finally we test the reliability of our method by applying it to controlled samples of segment Cox process and find that our method successfully recovers the length of the inputted segments. Summarizing these results we conclude that the statistically significant length scale of filaments is a robust measure of the galaxy distribution. ", "introduction": "The fact that the galaxies appear to be distributed along filaments which are interconnected to form a web like structure which is often reffered as the `cosmic web' is one of the most striking visual feature in all the present and past redshift surveys (e.g. , CfA , \\citealt{gel}; LCRS, \\citealt{shect}; 2dFGRS, \\citealt{colles} and SDSS, \\citealt{stout}). The analysis of filamentary patterns in the galaxy distribution has a long history dating back to a few papers in the late-seventies and mid-eighties by \\citet{joe}, \\citet{einas4}, \\citet{zel}, \\citet{shand1} and \\citet{einas1}. Filaments are the most striking visible patterns seen in the galaxy distribution (e.g. \\citealt{gel}, \\citealt{shect}, \\citealt{shand2}, \\citealt{bharad1}, \\citealt{mul}, \\citealt{basil}, \\citealt{doro2}, \\citealt{pimb}, \\citealt{pimb1}, \\citealt{pandey}). A review on a number of physically motivated and statistical methods to define filaments is provided in \\citet{pimb2}. The percolation analysis (eg. \\citealt{shand1}, \\citealt{einas1}), the genus statistics (eg. \\citealt{gott}), the minimal spanning tree (e.g. \\citealt{barrow}), the Voronoi tessellation (\\citealt{ike}, \\citealt{weygaert}), the Minkowski functionals (eg. \\citealt{mecke}, \\citealt{smal}) and the `Shapefinders' \\citep{sahni} are some of the useful statistical tools introduced to quantify the topology and geometry of the galaxy distribution. \\citet{stoi1} propose to apply a marked point process to automatically delineate filaments in the galaxy distribution. \\citet{colberg} studied the intercluster filamentary network in high resolution N-body simulations of structure formation in a $\\Lambda$CDM Universe. \\citet{arag} use Multiscale Morphology Filter technique to identify wall-like and filament-like structures in cosmological N-body simulations. \\citet{sus} propose a skeleton formalism to quantify the filamentary structure in three dimensional density fields. \\citet{stoi2} propose to apply an object point process to objectively identify filaments in galaxy redshift surveys. \\citet{sarkar1} propose the Local Dimension to locally quantify the shape of large scale structures in the neighbourhood of different galaxies in the Cosmic Web. The SDSS \\citep{york} is currently the largest galaxy redshift survey. In an earlier work \\citep{pandey} (hereafter Paper I) we have analysed the filamentarity in the equatorial strips of this survey. These strips are nearly two dimensional and we have projected the data onto a plane and analysed the resulting 2-D galaxy distribution. We find evidence for connectivity and filamentarity in excess of that of a random point distribution, indicating the existence of an interconnected network of filaments. We find that filaments are statistically significant upto length scales $80 \\, h^{-1} {\\rm Mpc}$ and not beyond \\citep{pandey}. All the structures spanning length-scales larger than this length scale are the result of chance alignments. This is consistent with an earlier analysis by \\citet{bharad2} where they show that in Las Campanas Redshift Survey (LCRS) the largest length-scale at which filaments are statistically significant is between $70$ to $80 \\, h^{-1}$Mpc. Further we show that the average filamentarity of the galaxy distribution depends on various physical properties of galaxies such as luminosity, colour, morphology and star formation rate \\citep{pandey1, pandey3}. It would be interesting to know if the statistically significant length scale of filaments also depends on different galaxy properties. In the present work we study if the statistically significant length scale of filaments depends on luminosity, colour and morphology of galaxies. Further it is also possible to measure the statistically significant length scale of filaments in mock galaxy distribution extracted from N-body simulation. The \\lcdm model is currently believed to be the minimal model which is consistent with most cosmological data (\\citealt{efst}; \\citealt{perci}; \\citealt{teg1}; \\citealt{sperg}; \\citealt{sperg1}; \\citealt{komat}). It would be interesting to measure the statistically significant length scale of filaments in N-body simulations of $\\Lambda$CDM model and compare it with measured values in the SDSS \\citep{pandey}. The N-body simulations primarily predict the clustering of the dark matter. This should be contrasted with the fact surveys reveal only the bright side of the matter distribution. Galaxy formation is a complicated process and the exact relation between the distribution of the galaxies and the dark matter is not well understood. It is now generally accepted that the galaxies are a biased tracer of the dark matter distribution (e.g., \\citealt{kais}; \\citealt{mo}) and on large scales one expects the fluctuations in the galaxy and the dark matter distribution to be linearly related through the linear bias parameter $b$. Mock galaxy distributions with different bias can be simulated following this assumption. The average filamentarity of the simulated galaxy distribution is found to be very sensitive to the bias parameter \\citep{bharad3, pandey2}. It would be interesting to know how the statistically significant length scale of filaments depends on the bias parameter. In the present work we test if the statistically significant length scale of filaments depends on bias. Earlier we find that the average filamentarity is sensitive to the area and galaxy number density of the samples \\citep{pandey1}(hereafter Paper II) and this statistics can be used for a meaningful comparison between two different galaxy samples only when they have the same volume (identical shape and size) and galaxy number density. In the present work we would also like to test if the statistically significant length scale of filaments depends on area and galaxy number density of the samples. Finally we simulate segment Cox process \\citep{pons} with different segment lengths and test the efficiency of our method in measuring the statistically significant length scale of filaments in the mock samples drawn from these simulations. A brief outline of our paper follows. Section 2 describes the data and method of analysis, our results and conclusions are presented in Section 3. ", "conclusions": "We use ``Shuffle'' to determine the statistically significant length scale of filaments in the seven SDSS strips in magnitude bin 1 and bin 2 (Table~1 and Table~2). We label bin 1 as Faint and bin 2 as Bright. The top two panels of Figure \\ref{fig:2} shows the results of ``Shuffle'' on strip 1 from bin 1 and bin 2 respectively. In top two panels of Figure \\ref{fig:2} we see that shuffling the data with $L=10$ causes a large drop in the average filamentarity ($F_2$). The average filamentarity increases as the value of $L$ increases and the value of average filamentarity of the unshuffled slice comes within $1 -\\sigma$ error bars when the data is shuffled with $L=90$ and $L=70$ for Bright and Faint galaxies respectively. Difference between the average filamentarity of shuffled and unshuffled slice is quantified by $\\chi^2/\\nu$ and the top two panels of Figure \\ref{fig:3} shows $\\chi^2/\\nu$ as a function of $L$. In top two panels of Figure \\ref{fig:3} we see that $\\chi^2/\\nu$ approaches 1 at $L=90$ for Bright and $L=70$ for Faint galaxies. So in strip 1 from bin1 and bin 2 the filaments are statistically significant upto length scales $80 \\, h^{-1} {\\rm Mpc}$ and $60 \\, h^{-1} {\\rm Mpc}$ respectively. The results from the other strips are not shown here. Averaging the results from all the nine strips we find that the filaments are statistically significant upto length scales $84 \\pm 16 \\, h^{-1} {\\rm Mpc}$ and $71 \\pm 16 \\, h^{-1} {\\rm Mpc}$ in bin 1 and bin 2 respectively. The right and left middle panels of Figure \\ref{fig:2} shows the results of ``Shuffle'' for Red and Blue galaxies (Table~2) from strip 1 in bin 2. The right and left middle panels of Figure \\ref{fig:2} shows that a large drop in the average filamentarity is observed when the data is shuffled with $L=10$. In both the panels we see that shuffling the data with $L=80$ increase the average filamentarity and brings it back in agreement with the actual data. In two middle panels of Figure \\ref{fig:3} we see that in both class of galaxies $\\chi^2/\\nu$ approaches 1 at $L=80$ establishing that for both Red and Blue galaxies the filaments are statistically significant upto length scales $70 \\, h^{-1} {\\rm Mpc}$. Here we have not shown the results from the other strips. Combining results from all the nine strips we find that the filaments are statistically significant upto length scales $77 \\pm 10 \\, h^{-1} {\\rm Mpc}$ and $71 \\pm 10 \\, h^{-1} {\\rm Mpc}$ for Red and Blue galaxies respectively. We show the results of ``Shuffle'' for the Early and Late type galaxies (Table~2) in the right and left bottom panels of Figure \\ref{fig:2}. Here also we see that the average filamentarity decreases when the data is shuffled with $L=10$ and it increases and matches with the actual data when the data is shuffled with $L=80$. The $\\chi^2/\\nu$ approaches 1 at $L=80$ for both Early and Late type galaxies as shown in two bottom panels of Figure \\ref{fig:3}. This establishes that for both Early and Late type galaxies the filaments are statistically significant upto length scales $70 \\, h^{-1} {\\rm Mpc}$ and not beyond. We have shown here only the results for strip 1 in bin 2. Combining the results from all the strips the filaments are found to be statistically significant upto length scales $75 \\pm 10 \\, h^{-1} {\\rm Mpc}$ and $74 \\pm 17 \\, h^{-1} {\\rm Mpc}$ for Early and Late type galaxies respectively. It is to be noted in middle and bottom two panels of Figure \\ref{fig:2} that the Red and Early type galaxies have higher degree of filamentarity as compared to Blue and Late type galaxies at smaller values of filling factor. We reported this effect in \\citet{pandey1}. It is interesting to note that although the average filamentarity ($F_2$) depends on luminosity, colour and morphology of galaxies, the statistically significant length scale of filaments does not depend on these galaxy properties. In Figure \\ref{fig:4} we show the results of ``Shuffle'' in simulated SDSS NGP strips with different biases. The top left panel show the result for the dark matter distribution from $\\Lambda$CDM N-body simulations. In this case galaxies are assumed to exactly trace the dark matter and the bias parameter $b=1$. We see that shuffling the simulated data with $L=10$ causes a large drop in the average filamentarity of the simulated galaxy distribution. With increasing L the value of average filamentarity of the shuffled slice slowly approaches the values corresponding to the unshuflled data. It is found that at $L=90$ the two are within $1 -\\sigma$ error bars. The top right panel shows the result for b=1.2. We see a very similar result and the unshuflled and shuffled data agrees well when the data is shuffled with $L=90$. In top left and right panels of Figure \\ref{fig:5} we see that $\\chi^2/\\nu$ approaches 1 at $L=90$ for both $b=1$ and $b=1.2$ and hence filaments are statistically significant upto length scales $80 \\, h^{-1} {\\rm Mpc}$ for both cases. The results are shown for one single mock SDSS strip. Averaging the results from all the nine strips it is found that for b=1 and b=1.2 filaments are statistically significant upto length scales $93 \\pm 16 \\, h^{-1} {\\rm Mpc}$ and $92 \\pm 15 \\, h^{-1} {\\rm Mpc}$ respectively. The bottom left panel of Figure \\ref{fig:4} shows the result of shuffling on simulated SDSS strips for a bias b=0.8. As usual shuffling the data with $L=10$ causes a drop in average filamentarity of the shuffled slice but the drop is smaller as compared to other bias values. As L increases the average filamentarity slowly grows and approaches the values corresponding to the unshuffled data at $L=90$. In the bottom right panel of Figure \\ref{fig:4} results are shown for a high bias b=1.5. Shuffling shows similar results for b=1.5 but a relatively large drop is seen when the data is shuffled with $L=10$. With increasing L the average filamentarity grows relatively faster as compared to b=0.8 and finally the average filamentarity in the shuffled data levels up with the unshuffled data when $L=80$ is used for shuffling. In the bottom left and right panels of Figure \\ref{fig:5} it is found that $\\chi^2/\\nu$ approaches 1 at $L=90$ and $L=80$ indicating that filaments are statistically significant upto length scales $80 \\, h^{-1} {\\rm Mpc}$ and $70 \\, h^{-1} {\\rm Mpc}$ for b=0.8 and b=1.5 respectively. These are the results shown for a single mock SDSS strip and by combining the results from the other strips we get that the filaments are statistically significant upto length scales $91 \\pm 13 \\, h^{-1} {\\rm Mpc}$ and $90 \\pm 15 \\, h^{-1} {\\rm Mpc}$ for $b=0.8$ and $b=1.5$ respectively. \\begin{figure} \\rotatebox{-90}{\\scalebox{.4}{\\includegraphics{plot5.ps}}} \\caption{This shows $\\chi^2/\\nu$ at different shuffling lengths for a simulated SDSS NGP strip with different bias values as indicated in each panel. The black dotted line indicates $\\chi^2/\\nu=1$.} \\label{fig:5} \\end{figure} \\begin{figure} \\rotatebox{-90}{\\scalebox{.4}{\\includegraphics{plot6.ps}}} \\caption{This shows the statistically significant length scale of filaments as a function of bias b. The error bars are $1-\\sigma$ error bars measured from nine mock galaxy catalogues.} \\label{fig:6} \\end{figure} \\begin{figure} \\rotatebox{-90}{\\scalebox{.4}{\\includegraphics{plot7.ps}}} \\caption{This shows the statistically significant length scale of filaments as a function area of the samples described in Table~3. The error bars are $1 - \\sigma$ error bars measured from nine mock galaxy catalogues. } \\label{fig:7} \\end{figure} \\begin{figure} \\rotatebox{-90}{\\scalebox{.4}{\\includegraphics{plot8.ps}}} \\caption{This shows the statistically significant length scale of filaments as a function galaxy number density of the samples described in Table~3. The error bars are $1 - \\sigma$ error bars measured from nine mock galaxy catalogues. } \\label{fig:8} \\end{figure} It is to be noted in Figure \\ref{fig:4} that as the bias $b$ is increased the average filamentarity ($F_2$) increases at smaller $FF$. We reported this effect in \\citet{bharad3}. But interestingly the shuffling length at which the avergae filamentarity of the unshuflled and shuffled data agrees does not depend on bias. Figure \\ref{fig:6} shows the statistically significant length scale of filaments as a function of bias. This Figure shows that the statistically significant length scale of filaments is nearly independent of bias. The mean value is $~90 \\, h^{-1} {\\rm Mpc} $ for the bias values $b=0.8-1.5$. It tends to decrease for a high value of bias $b=1.8$. This could be possibly related to the fact that at very high values of bias we preferentially identify only very high density regions which are mostly related to cluster like structures rather than filaments. Earlier in Paper I we established that the filaments are statistically significant upto length scales of $80 \\, h^{-1} {\\rm Mpc}$. This measured value lies well within $1-\\sigma$ error bars of the measured values for mock galaxy distribution from $\\Lambda$CDM N-body simulations. We conclude that the $\\Lambda$CDM is consistent with SDSS observations. The galaxy samples constructed over different magnitude ranges have different area and galaxy number density (Table~2). Earlier analysis in Paper II shows that the average filamentarity of the galaxy distribution depends on the area and galaxy number density of the galaxy samples and filamentarity between different galaxy samples can only be compared when the samples have same area, geometry and number density of galaxies. So it is important to know if the statistically significant length scale of filaments depends on these factors before we can compare it's measured values among various galaxy samples having different areas and galaxy number densities. We prepare two different set of mock galaxy samples namely area 1 and area 2 (Table~3) which have the same galaxy number density as the SDSS NGP strip (area 3) but cover different redshift ranges and hence have different areas as listed in Table~3. We extract three mock galaxy samples from each of the $\\Lambda$CDM dark matter simulation for each set of galaxy samples area 1 and area 2. We determine the statistically significant length scale of filaments for these galaxy samples. The analysis for the mock SDSS NGP strip which we name as area 3 in Table~3 are already done. We plot the statistically significant length scale of filaments as a function of the area of the galaxy samples in Figure \\ref{fig:7}. From this figure we see that the samples of smaller area tend to have a smaller length scale of filaments. But this trend is weak and the size of the error bars are quite large. We conclude that the statistically significant length scale of filaments are nearly independent of the area of the samples. We next prepare two different set of mock galaxy samples namely density 1 and density 3 (Table~3) which have the same area as the SDSS NGP strip (density 2) but have different number densities listed in Table~3. We extract three mock galaxy samples from each of the $\\Lambda$CDM dark matter simulation for each set of galaxy samples density 1 and density 3. We determine the statistically significant length scale of filaments for these galaxy samples. The analysis for the mock SDSS NGP strip which we name as density 2 in Table~3 are already done. We plot the statistically significant length scale of filaments as a function of the number density of galaxies of the samples in Figure \\ref{fig:8}. We see a trend of low density samples to have a smaller length scale of filaments. But the error bars are quite large and the dependence is weak. We conclude that the statistically significant length scale of filaments weakly depends on the number density of galaxies. In a recent work with SDSS LRG samples \\citep{pandey4} having much larger area and lower density we find that the statistically significant length scale of filaments does not change much as compared to it's value measured in SDSS Main galaxy samples. Thus the statistically significant length scale of filaments is a more robust statistics than the average filamentarity. It does not depend on properties (size and number density) of galaxy samples , and physical properties of galaxies. It emerges as a robust measure of the large scale structures. The statistically significant length scale of filaments could be also thought of as an indicator of the scale beyond which the galaxy distribution become homogeneous because it tells us the length scale beyond which no coherent structures exist in the galaxy distribution. Multifractal analysis in SDSS DR1 \\citep{yadav} and recently in SDSS DR6 \\citep{sarkar} show that the galaxy distribution becomes homogeneous at a length scale between $60$ to $70 h^{-1} \\, {\\rm Mpc}$. \\citet{yadav} considered the transition to homogeneity in $\\Lambda$CDM simulations with different bias $b=1,1.6,2$ and find that irrespective of the bias values the distribution become homogeneous at a length scale between $60$ to $70 h^{-1} \\, {\\rm Mpc}$. These results are consistent with our findings. Finally we test the reliability of our method by applying it to different controlled samples (Table~4) of 3D segment Cox process. We use segments of two different length $l=20\\, h^{-1} {\\rm Mpc}$ and $l=80\\, h^{-1} {\\rm Mpc}$ for simulating the 3D segment Cox process. Two different sets of 2D mock samples drawn from these 3D simulations (Table~4) are separately analyzed. The results are shown in top two panels of Figure \\ref{fig:9} and Figure \\ref{fig:10}. We see in top left panel of Figure \\ref{fig:9} that filamentarity decrease by a very small amount when the mock samples having $l=20\\, h^{-1} {\\rm Mpc}$ are shuffled with $L=10$ and there are virtually no drop in filamentarity when the data is shuffled with $L=20$. The top right panel of the same figure shows the result of ``Shuffle'' on cotrolled samples of segment Cox process with segment length $l=80\\, h^{-1} {\\rm Mpc}$. We see a significant drop in the average filamentarity when the data is shuffled with $L=10$ and the average filamentarity finally saturates to the values corresponding to the unshuffled data when $L=60$ is used for shuffling. The $\\chi^2/\\nu$ as a function of $L$ for these two cases are shown in top two panels of Figure \\ref{fig:10}. We see that $\\chi^2/\\nu$ approaches 1 at $L=20$ and $L=60$ in these two cases respectively. This indicates that for these 2D mock samples having $l=20\\, h^{-1} {\\rm Mpc}$ and $l=80\\, h^{-1} {\\rm Mpc}$, the filaments are found to be statistically significant upto length scales $10 \\, h^{-1} {\\rm Mpc}$ and $50 \\, h^{-1} {\\rm Mpc}$ respectively. Here the results are shown only for a single simulated mock sample. By averaging the results from all the nine simulated mock samples we find that in the above two cases the filaments are statistically significant upto length scales $18 \\pm 7 \\, h^{-1} {\\rm Mpc}$ and $52 \\pm 11 \\, h^{-1} {\\rm Mpc}$ respectively. We note that the values of statistically significant length scale of filaments recovered by our method for straight segments simulated here are a little shorter but close to the original values of $l$ used in the simulations. The simulations are carried out in 3D boxes and our mock samples are the 2D projections of nearly two dimensional samples drawn from these boxes. This possibly destroys and shortens a lot of segments and the effects are expected to be more prominent for longer segments as reflected in our results. This effect indicates that also for the 2D real galaxy samples from SDSS which are constructed out of a 3D galaxy distribution, Shuffle possibly underestimates the values of statistically significant length scale of the real 3D filaments. The effects of slicing and projections can be avoided if we simulate the segment Cox process in 2D as in this case the segments lie on the same plane (Figure \\ref{fig:11}). We apply Shuffle to a set of mock 2D strips constructed out of simulations of 2D segment Cox process (Table~5). We use two different segment lengths $l=20\\, h^{-1} {\\rm Mpc}$ and $l=80\\, h^{-1} {\\rm Mpc}$ to simulate the 2D segment Cox process. We use the same segment lengths as used in the simulations of 3D segment Cox process so as to test the effect of slicing and projection. We show the results in bottom two panels of Figure \\ref{fig:9} and Figure \\ref{fig:10}. Bottom left panel of Figure \\ref{fig:9} shows that when the mock samples with $l=20\\, h^{-1} {\\rm Mpc}$ are shuffled with $L=10$ the average filamentarity of the shuffled data falls below that of the unshuffled data. The average filamentarity of the shuffled data saturates to the unshuffled data when the data is shuffled with $L=40$. The bottom right panel of Figure \\ref{fig:9} shows a significant drop in the average filamentarity when the mock samples with $l=80\\, h^{-1} {\\rm Mpc}$ are shuffled with $L=10$. The average filamentarity of the shuffled data increases slowly with increasing shuffling length and finally saturates to the unshuffled data at shuffling length $L=90$. We see in bottom two panels of Figure \\ref{fig:10} the $\\chi^2/\\nu$ as a function of $L$ approaches 1 at $L=40$ and $L=90$ in these two cases respectively. This indicates that for these controlled samples of 2D segment Cox process having $l=20\\, h^{-1} {\\rm Mpc}$ and $l=80\\, h^{-1} {\\rm Mpc}$, the filaments are statistically significant upto length scales $30 \\, h^{-1} {\\rm Mpc}$ and $80 \\, h^{-1} {\\rm Mpc}$ respectively. We have shown the results for a single simulated strip in these plots. We combine the results from all the nine simulated strips and find that the filaments are statistically significant upto length scales $25 \\pm 8 \\, h^{-1} {\\rm Mpc}$ and $82 \\pm 24 \\, h^{-1} {\\rm Mpc}$ for simulations of 2D segment Cox process with segment length $l=20\\, h^{-1} {\\rm Mpc}$ and $l=80\\, h^{-1} {\\rm Mpc}$ respectively. We note that our method can successfully recover better the length of inputted segments when the mock samples are drawn from simulations of 2D segment Cox process where the effects of slicing and projections are absent. There would be another effect due to the boundaries of the samples which again shortens the length of some segments near the boundaries but this effect is not so important in our analysis as the extent of the samples are much larger than the length of the inputted segments. This analysis suggets that the actual length scale of the real 3D filaments can be recovered with Shuffle if the analysis is extended in 3D. We propose to take up this in a future work. In conclusion we note that although the average filamentarity of the galaxy distribution depends on different galaxy properties, the statistically significant length scale of filaments does not depend on luminosity, colour and morphology of galaxies. We find that the measured values of the statistically significant length scale of filaments in SDSS are consistent with that measured from $\\Lambda$CDM simulations. We considered the $\\Lambda$CDM model with different values of bias and find that the statistically significant length scale of filaments is nearly independent of bias. Different class of galaxies are differently biased with respect to underlying dark matter distribution and this result is possibly related to the fact that statistically significant length scale of filaments is nearly the same for different class of galaxies. Unlike the average filamentarity, the statistically significant length scale of filaments is nearly independent of the area and galaxy number density of the samples. This establishes the statistically significant length scale of filaments as a robust statistics of galaxy distribution. \\begin{figure} \\rotatebox{0}{\\scalebox{.6}{\\includegraphics{plot11.ps}}} \\caption{ This shows a simulation of 2D segment Cox process with two different segment length $l=20\\, h^{-1} {\\rm Mpc}$ and $l=80\\, h^{-1} {\\rm Mpc}$ (as indicated in each panel) over a region which has identical geometry to the SDSS NGP equatorial strip. These simulated strips have same values of $l$ and $\\lambda_{l}$ listed in Table~5 but different values of $\\lambda_{s}$ are chosen so as to lower the density of the segments and make them visibly distinguishable in this plot.} \\label{fig:11} \\end{figure} \\begin{figure} \\rotatebox{-90}{\\scalebox{.4}{\\includegraphics{plot9.ps}}} \\caption{ This shows the Average Filamentarity as a function of Filling Factor ($FF$) for a mock SDSS NGP strip constructed out of 3D simulation of segment Cox process (Table~4) and 2D simulation of segment Cox process (Table~5) together with the results for the shuffled data for two values of $L$ shown in the figure. The 2D simulations of segment Cox process contain all the segments in the same plane and there are no effects because of the slicing process or the projections which arises when samples are constructed out of 3D simulations of segment Cox process. We see in this plot that ``Shuffle\" is able to recover the length of the longer segments more efficiently when samples are constructed out of 2D simulations instead of 3D simulations of segment Cox process. } \\label{fig:9} \\end{figure} \\begin{figure} \\rotatebox{-90}{\\scalebox{.4}{\\includegraphics{plot10.ps}}} \\caption{ This shows $\\chi^2/\\nu$ at different shuffling lengths ($L$) for a mock SDSS NGP strip constructed out of 3D segment Cox process and 2D segment Cox process simulated with two different segment length $l=20\\, h^{-1} {\\rm Mpc}$ and $l=80\\, h^{-1} {\\rm Mpc}$. The black dotted line indicates $\\chi^2/\\nu=1$.} \\label{fig:10} \\end{figure}" }, "0910/0910.4579.txt": { "abstract": "{The dynamical ejection of stars from star clusters affects the shape of the stellar mass function (MF) in these clusters, because the escape probability of a star depends on its mass. This is found in $N$-body simulations and has been approximated in analytical cluster models by fitting the evolution of the MF. Both approaches are naturally restricted to the set of boundary conditions for which the simulations were performed.} {The objective of this paper is to provide and to apply a simple physical model for the evolution of the MF in star clusters for a large range of the parameter space. It should also offer a new perspective on the results from $N$-body simulations.} {A simple, physically self-contained model for the evolution of the stellar MF in star clusters is derived from the basic principles of two-body encounters and energy considerations. It is independent of the adopted mass loss rate or initial mass function (IMF), and contains stellar evolution, stellar remnant retention, dynamical dissolution in a tidal field, and mass segregation.} {The MF evolution in star clusters depends on the disruption time, remnant retention fraction, initial-final stellar mass relation, and IMF. Low-mass stars are preferentially ejected after $t\\sim 400$~Myr. Before that time, masses around 15---20\\% of the maximum stellar mass are lost due to their rapid two-body relaxation with the massive stars that still exist at young ages. The degree of low-mass star depletion grows for increasing disruption times, but can be quenched when a large fraction of massive remnants is retained. The highly depleted MFs of certain Galactic globular clusters are explained by the enhanced low-mass star depletion that occurs for low remnant retention fractions. Unless the retention fraction is exceptionally large, dynamical evolution always decreases the mass-to-light ratio. The retention of black holes reduces the fraction of the cluster mass in remnants because white dwarfs and neutron stars have masses that are efficiently ejected by black holes.} {The modeled evolution of the MF is consistent with $N$-body simulations when adopting identical boundary conditions. However, it is found that the results from $N$-body simulations only hold for their specific boundary conditions and should not be generalised to all clusters. It is concluded that the model provides an efficient method do understand the evolution of the stellar MF in star clusters under widely varying conditions.} % ", "introduction": "\\label{sec:intro} The evaporation of star clusters is known to change the shape of the underlying stellar mass function\\footnote{Hereafter, `mass function' is referred to as `MF'.} \\citep{henon69,chernoff90,vesperini97b,takahashi00,portegieszwart01,baumgardt03}. This phenomenon has been used to explain the observed MFs in globular clusters \\citep{richer91,demarchi07,demarchi07b}, which are flatter than typical initial mass functions \\citep[IMFs, e.g.][]{salpeter55,kroupa01}. In addition, the effect of a changing MF on cluster photometry has been investigated \\citep{lamers06,kruijssen08,anders09}. This has been shown to explain the low mass-to-light ratios of globular clusters \\citep{kruijssen08b,kruijssen09} and to have a pronounced effect on the inferred globular cluster mass function \\citep{kruijssen09b}. The existing parameterised cluster models that incorporate a description of low-mass star depletion are restricted by the physically self-contained models on which they are based. Some studies \\citep{lamers06,kruijssen08} assume an increasing lower stellar mass limit to account for the evolving MF, others \\citep{anders09} fit a changing MF slope to $N$-body simulations. In both cases, the models are accurate for a certain range of boundary conditions, but they do not include a physical model and are therefore lacking flexibility. While $N$-body simulations do include the appropriate physics, they are very time-consuming. As a result, only a limited number of clusters can be simulated and the applicability of the simulations is thus restricted to the specific set of boundary conditions for which they have been run. It would be desirable to obtain a simple physical model for the evolution of the MF, which would have a short runtime and could be used independently of $N$-body simulations. Forty years ago, a pioneering first approach to such a model was made by \\citet{henon69}, who considered the stellar mass-dependent escape rate of stars from star clusters. However, the applicability of his model was limited due to a number of assumptions that influenced the results. First of all, \\citet{henon69} assumed that the clusters exist in isolation and neglected the tidal field. As a consequence, the ejection of a star could only occur by a single, close encounter and the repeated effect of two-body relaxation was not included. Secondly, the {distribution of stars was independent of stellar mass, i.e.} mass segregation was not included. Both mass segregation and the influence of a tidal field are observed in real clusters, and can be expected to affect the evolution of the MF. The aim of this paper is to derive a physical description of the evolution of the stellar MF in star clusters, alleviating the assumptions that were made by \\citet{henon69}. This should explain the results found in $N$-body simulations and observations, while providing the required flexibility to explore the properties of star clusters with simple, physically self-contained models. The outline of this paper is as follows. In Sect.~\\ref{sec:model}, total mass evolution of star clusters is discussed. A recipe for the evolution of the MF is derived in Sect.~\\ref{sec:mf}, covering stellar evolution, the retain of stellar remnants, dynamical dissolution and mass segregation. The model is compared to $N$-body simulations in Sect.~\\ref{sec:comp}. In Sect.~\\ref{sec:results}, the model is applied to assess the evolution of the MF for different disruption times and remnant retention fractions. The consequences for other cluster properties are also considered. This paper is concluded with a discussion of the results and their implications. ", "conclusions": "\\label{sec:disc} The results of this paper show that the stellar MFs in star clusters differ strongly from their initial forms due to dynamical cluster evolution. The specific kinds of these differences depend on the properties of the star clusters and their tidal environment, most importantly on the disruption time, remnant retention fraction, and IMF.\\footnote{Although not specifically shown in this paper (but not surprisingly), the differences also depend on the initial-final stellar mass relation.} A physical model for the evolution of the stellar MF is presented in which two-body relaxation leads to a stellar mass dependence of the ejection rate. For any particular stellar mass, the ejection rate is determined by the typical proximity of that mass to the escape energy and by the timescale on which the two-body relaxation with the other stars takes place. Combined with a prescription for stellar evolution, stellar remnant production, and remnant retention using kick velocity dispersions, this provides a description for the total evolution of the MF. {\\it This description is independent of the adopted total mass evolution}. {The model shows that the slope of the mass function is a possible indicator for the mass fraction that has been lost due to dissolution, provided that the IMF does not vary and the remnant retention fraction has been fairly similar for young globular clusters.\\footnote{Any variability of the retention fraction would induce substantial scatter, see Sect.~\\ref{sec:remn} and Fig.~\\ref{fig:alphatdis}.}} For the exact same initial conditions, the model shows excellent agreement with $N$-body simulations of the evolving MF by \\citet{baumgardt03}. However, an important advantage of the presented model compared to the (more accurate) $N$-body simulations is its short runtime and corresponding flexibility. It can be easily applied to compute the evolution of clusters for a large range of initial conditions. The results can then be used to identify interesting cases for more detailed and less simplified calculations with $N$-body or Monte Carlo models. {The most important simplification of the model is neglecting the effect of binary encounters on the stellar mass dependence of the ejection rate. To incorporate binaries, a conclusive census of the binary population in star clusters would be required, which is not yet available. Nonetheless, it is possible to make a qualitative estimate for the effect of binaries. The encounter rate of binaries would typically be higher than that of individual stars, because the cross section of binaries is larger. This would increase the relative escape rate at the stellar mass for which the binary fraction\\footnote{The fraction of stars residing in binary or multiple systems.} peaks. This binary fraction is found to increase with primary mass \\citep[see e.g.][]{kouwenhoven09}. Because massive stars are removed by stellar evolution, this implies that the binary fraction decreases with age, which is in agreement with the low binary fraction observed in globular clusters \\citep[$\\sim 2\\%$, e.g.][]{richer04}. The effect of binaries on the evolution of the mass function would thus be most notable if the majority of the dynamical mass loss occurs at ages $< 50$~Myr (the typical lifetime of an $8~\\msun$ star), in which case it would somewhat enhance the relative escape rate of the most massive stars.} The model is applied to investigate the influence of the disruption time and remnant retention on the evolution of the MF and integrated photometric properties of star clusters. For total disruption times $\\tdis<1$~Gyr, the modeled relative ejection rate is highest at a certain `sweet spot' mass that is typically 15---20\\% of the mass of the most massive objects in the cluster. For longer lifetimes, the evolution of the MF is dominated by low-mass star depletion, unless the retention fraction of massive stellar remnants is larger than 0.25. Only in the particular case of such a high retention fraction, the $M/L$ ratio is increased by dynamical evolution when the cluster approaches total disruption. In all other scenarios, the $M/L$ ratio decreases because the most massive (luminous) stars are kept.\\footnote{This process differs from a possible variability of the proportionality between the velocity dispersion and the cluster mass, which concerns a much shorter timescale \\citep[e.g.][]{boily09}.} When defining the slope of the MF in the range 30---80\\% of the maximum stellar mass, this gives a clear relation between the MF slope and the $M/L$ ratio. For slopes that are defined in fixed mass ranges, there is not necessarily a correlation between slope and $M/L$ ratio if $\\tdis<1$~Gyr. In clusters with a longer total disruption time, both quantities are related. Dynamical cluster evolution is found to induce some reddening of the integrated cluster colours, amounting up to 0.1---0.2~mag in $V-I$ for total disruption times $\\tdis<1.5$~Gyr. The fraction of the cluster mass that is constituted by remnants surprisingly decreases if more black holes are retained, because the black holes preferentially eject bodies around the masses of white dwarfs and neutron stars, which contain most of the total remnant mass. Contrary to what is suggested by other studies \\citep[e.g.][]{baumgardt03,anders09}, the evolution of the MF is not homologous. The reason that these studies concluded that its evolution is very similar for all clusters (also see Figs.~\\ref{fig:comp} and~\\ref{fig:comp2}), is that they assumed that all remnants were retained. It is illustrated in Fig.~\\ref{fig:mf2} that the differences between clusters with dissimilar disruption times disappear when the retention fraction increases. For realistic retention fractions, differences do arise. If two clusters with different initial masses have the same total disruption time, their MF evolution will be dissimilar due to their different remnant retention fractions and the impact of the retained remnants on the dynamical cluster evolution. Alternatively, if two clusters have equal initial masses but different total disruption times, for instance due to differences in their galactic location or environment, their MF evolution will be dissimilar due to the dynamical impact of the evolution of the maximum stellar mass. The larger variation of MF evolution that is found with presented model may also be able to explain observations of globular clusters in which the MF cannot be characterised by a single power law \\citep{demarchi00}. If the evolution of the MF were homologous, these features would likely be primordial \\citep{baumgardt03}, but this is not necessarily the case when using realistic retention fractions. Most other differences between the results presented in Sect.~\\ref{sec:results} and those from \\citet{baumgardt03} are also due to their assumption of full remnant retention. For example, their $M/L$ ratio evolution shows a smaller decrease than in Fig.~\\ref{fig:ml1}. This is explained in Fig.~\\ref{fig:ml2}, where it is shown that dynamical evolution reduces the $M/L$ ratio by a smaller amount if the retention fraction is larger. {Studies on the fractal nature of cluster formation show that star clusters are initially substructured \\citep{elmegreen00,bonnell03}. Even though this substructure is typically erased on a crossing time, it can induce primordial mass segregation in star clusters \\citep{mcmillan07,allison09}.} The influence of primordial mass segregation on the evolution of the MF has recently been investigated by \\citet{baumgardt08b} and \\citet{vesperini09}. While \\citet{baumgardt08b} do not include stellar evolution and concentrate on two-body relaxation, \\citet{vesperini09} do include stellar evolution. They show that for some degrees of primordial mass segregation, the mass loss by stellar evolution can induce additional dynamical mass loss that strongly decreases the total disruption time. For clusters that survive for a Hubble time, the MF evolution in the case of primordial mass segregation is very similar to an initially unsegregated cluster. \\citet{vesperini09} conclude that the evolution of the MF is only affected by primordial mass segregation for clusters in which the total disruption time is sufficiently decreased by the induced mass loss. In that case, the slope of the MF remains much closer to its initial value than it would in clusters without primordial mass segregation. Their conclusion is consistent with the model presented in this paper, because the evolution of the MF is determined by the most massive stars at the time when the largest mass loss occurs (see Figs.~\\ref{fig:chinorm} and~\\ref{fig:mf1}). This induced mass loss enters the model in terms of the absolute mass loss rate in Eq.~\\ref{eq:dmdt}, not in the stellar mass-dependent escape rate per unit mass loss rate of Eq.~\\ref{eq:chi}. A change in total mass loss rate is not the only consequence of primordial mass segregation. \\citet{baumgardt08b} have shown that low-mass star depletion is enhanced for clusters without stellar evolution that are primordially mass-segregated. This occurs because energy equipartition is reached on a shorter timescale and because of their use of a fixed ($m_{\\rm max}=1.2~\\msun$) maximum stellar mass. As a result, there are no massive bodies to increase the ejection rate of intermediate mass stars (see Fig.~\\ref{fig:maxchinorm}), implying that only the low-mass stars are preferentially lost. In the present paper, mass segregation is assumed to arise dynamically, but the model could in principle be adapted to cover primordial mass segregation by setting $c_2=0$ and adjusting $c_1$ to the initial velocity distribution until it is erased by dynamical evolution (see Eq.~\\ref{eq:tseg}), after which the values from Sect.~\\ref{sec:mf} can be used.\\footnote{As explained in Sect~\\ref{sec:mf}, $c_1$ represents the ratio of the mean speed squared to the central escape velocity squared that depends on the degree of mass segregation (and thus on the IMF). On the other hand, $c_2$ is a proportionality constant in the expression for the onset of the stellar mass-dependent ejection of stars, which depends on the concentration or King parameter.} This does not necessarily yield enhanced low-mass star depletion for clusters with a complete IMF (including masses $m>1.2~\\msun$) because of the presence of massive stars or remnants. \\begin{figure}[t] \\resizebox{\\hsize}{!}{\\includegraphics{alphatdis.ps}} \\caption[]{\\label{fig:alphatdis} MF slope versus remaining lifetime (assuming a globular cluster age of 12~Gyr). Diamonds represent the observed values from \\citet{demarchi07}, with typical errors as shown by the error bar in the lower right corner. The remaining lifetimes are taken from \\citet{baumgardt08b}. Dotted curves represent the model evolutionary tracks of clusters with $\\log{(M_{\\rm i}/\\msun)}=\\{6,6.25,6.5,6.75,7\\}$ from Sect.~\\ref{sec:remn} with $\\{\\sigma_{\\rm kick,wd},\\sigma_{\\rm kick,ns},\\sigma_{\\rm kick,bh}\\}=\\{4,100,200\\}$~km~s$^{-1}$, corresponding to $\\{f_{\\rm ret,wd},f_{\\rm ret,ns},f_{\\rm ret,bh}\\}=\\{0.983,0.022,0.003\\}$ for a $10^6~\\msun$ cluster. The solid line connects the present-day locations of the modeled clusters in the diagram (crosses), while the dashed line represents the same relation for $\\sigma_{\\rm kick,bh}=40$~km~s$^{-1}$ ($f_{\\rm ret,bh}=0.219$ for a $10^6~\\msun$ cluster). The dash-dotted line shows the homologous cluster evolution from \\citet{baumgardt03}. } \\end{figure} The presented model can be applied to the MFs of Galactic globular clusters that are observed by \\citet{demarchi07}. These MFs are more strongly depleted than is found in the $N$-body simulations by \\citet{baumgardt03}, which has been attributed to primordial mass segregation \\citep{baumgardt08b}. However, the observations can also very accurately be explained with the realistic remnant retention fractions that are used in the present paper. This is shown in Fig.~\\ref{fig:alphatdis}, where the observed MF slopes and remaining lifetimes of the globular clusters from \\citet{demarchi07} are compared with the globular cluster-like models from Sect.~\\ref{sec:remn} {($t_0=1$~Myr)}. The models are in much better agreement with the data than the $N$-body runs with complete remnant retention from \\citet{baumgardt03}. Deviations to other values of $\\alpha$ can occur due to variations in disruption time and remnant retention fractions, as is also shown in Fig.~\\ref{fig:alphatdis}. For example, a variation of the remnant kick velocity with metallicity in combination with the known variation of the disruption time \\citep[see, e.g.][]{kruijssen09,kruijssen09b} should be sufficient to cover the observed scatter. {The above line of reasoning provides an explanation for the the depleted MFs in Fig.~\\ref{fig:alphatdis} that is consistent with the simulations by \\citet{vesperini09}, who showed that the effects of primordial mass segregation are in fact suppressed in long-lived clusters due to the expansion caused by stellar evolution. This increases the relaxation time and yields an evolution of the MF that is very similar to the initially unsegregated scenario, indicating that primordial mass segregation is not a likely explanation for strongly depleted MFs.} Observations of the remnant composition of these globular clusters could reveal {a definitive answer as to} whether the depleted MFs are explained by primordial mass segregation or by dynamical evolution with a realistic remnant retention fraction. Dynamical cluster evolution does not appear to have a large effect on the colours of old (globular) clusters. The only way in which the colours could be affected beyond typical observational errors, is if globular clusters have lost substantial fractions of their masses during the first $\\sim$~Gyr after their formation. In that case, the dynamical evolution of the stellar MF in globular clusters may have implications for studies of colour bimodality \\citep[e.g.][]{larsen01} or the blue tilt \\citep[e.g.][]{harris06}. It could then also possibly explain the trend of increasing $V-K$ colour with decreasing $M/L_V$ ratio found by \\citet{strader09} for globular clusters in M31, because quickly dissolving clusters generally become redder and have reduced $M/L$ ratios. More research is needed to determine the role of the changing MF in the above properties of globular cluster systems. It can be concluded that the evolution of the stellar MF in star clusters is not as similar for all clusters as previously thought. Its precise evolution is determined by cluster characteristics like the disruption time, the remnant retention fraction, initial-final stellar mass relation, and the IMF. In order to decipher the evolution of observed star clusters, it is essential to record these characteristics and to relate them to possible scenarios for the internal evolution of clusters. That way, observables like the slope of the MF, the $M/L$ ratio, the broadband colours, and the mass fraction in remnants can be better understood." }, "0910/0910.5134_arXiv.txt": { "abstract": "We propose a new scenario for the early universe where there is a smooth transition between an early de Sitter-like phase and a radiation dominated era. In this model, the matter content is modelled by a new type of generalised Chaplygin gas \\cite{chaplygin} for the early universe, with an underlying scalar field description. We study the gravitational waves generated by the quantum fluctuations. In particular, we calculate the gravitational wave power spectrum, as it would be measured today, following the method of the Bogoliubov coefficients. We show that the high frequencies region of the spectrum depends strongly on one of the parameters of the model. On the other hand, we use the number of e-folds, along with the power spectra and spectral index of the scalar perturbations, to constrain the model observationally. ", "introduction": "\\label{sec1} The inflationary paradigm is the most consistent scenario to explain the origin of the large scale structure (LSS) of the universe, as well as the anisotropies in the cosmic microwave background radiation (CMBR) \\cite{LiddleLyth,Kinney:2009vz}. Originally, an early inflationary era in the universe was invoked as a mechanism to solve several shortcomings of the big bang theory \\cite{inflation}. Afterwards, it was realised that an early inflationary era of the universe generates density perturbations as seeds for the structure we see nowadays. In addition, a background of stochastic primordial gravitational waves (GW) is also predicted and it is originated from the vacuum fluctuations. The latter (GW) if ever detected in the future would provide an amusing source of information about the early universe. On the other hand, and most importantly, the predictions of the inflationary theory have been corroborated by several cosmological observations like the recent WMAP5 data \\cite{WMAP}. In this letter we propose a phenomenological model for the inflationary era and the subsequent radiation dominated epoch of the universe. In particular, we suggest a way to extend the framework of the generalised Chaplygin gas (GCG) to the first stages of the universe\\footnote{An alternative inflationary model inspired on the Chaplygin gas was proposed in Ref.~\\cite{Bertolami:2006zg} where only the inflationary era is accounted for.}. The GCG has attracted a lot of attention over the last years \\cite{chaplygin,chaplygin2,gorini,Barreiro:2008pn,GCGsingularity}. It was initially introduced in cosmology as a mean to unify the dark sectors of the universe, i.e. dark energy and dark matter, and has been contrasted with several cosmological observations \\cite{Barreiro:2008pn}. On the other hand, it has been as well shown that some GCG models can be an excellent framework to analyse dark energy related singularities \\cite{chaplygin2,GCGsingularity}. In this work, we investigate the possible imprints in the power spectrum of the stochastic background of GW, for the model we propose for the transition from the inflationary era to the radiation dominated epoch\\footnote{In Ref.~\\cite{{Fabris:2004dp}}, the spectrum of GW for a universe whose dark energy component corresponds to a Chaplygin gas was analysed. This scenario is different from ours.}. This transition is far from being well known and it can be of some interest to explore the signatures associated with it, shedding light on the inflationary model behind the accelerated expansion in the early universe. We calculate the GW production using the Bogoliubov coefficients which obey a set of differential equations \\cite{parker,Starobinsky,allen,mendes-henriques-moorhouse}. This method has advantages over the frequently used sudden transition approximation, because, associated with it, there is always an overproduction of gravitons of large frequencies, which is avoided in a natural way by the use of the continuous Bogoliubov coefficients \\cite{mendes-henriques-moorhouse}. This is also a very practical method to calculate the full spectrum, from the very low frequencies corresponding to the present cosmological horizon $10^{-17}$Hz, till those large, kHz up to GHz, frequencies associated with the transition between the inflationary and the radiation-dominated universe. The present letter is organised as follows. In section \\ref{sec2} we present our model based on a \\textit{new type of generalised Chaplygin gas for the early universe}. This fluid can be as well represented by a minimally coupled scalar field as shown in section \\ref{sec3}. In section \\ref{sec4}, we constrain our model observationally. In section \\ref{sec5}, we summarise the methodology used to obtain the spectrum of the stochastic GW which is based on the Bogoliubov coefficients. We present in section \\ref{sec:NumericalSimulations} the spectrum of the GW for the model introduced in sections \\ref{sec2} and \\ref{sec3}. Finally, in section \\ref{sec7} we present our conclusions. ", "conclusions": "\\label{sec7} \\label{sec:Conclusions} In this work, we investigate the production of GWs in a new modified generalised Chaplygin gas for the early universe (cf. Secs \\ref{sec2} and \\ref{sec3}). Through recent measurements of CMBR/LSS, the parameters of the model have been constrained. We have used the method of the continuous Bogoliubov coefficients to calculate the GW energy spectrum for different scales of energy and values of the $\\alpha$-parameter of the model. Besides the fact that the model suffers a small deviation from the preferred values of the spectral index $n_s$, the obtained spectra reveal a consistent picture corresponding to a smooth transition between the inflationary and the radiation dominated epochs of the universe, and constitute a significant imprint of this modified generalised Chaplygin gas. Finally, the strong variation of the high frequency range, is directly related with the decrease in the maximum of the potential $a^{\\prime\\prime}/a$ (see Eq.~(\\ref{6})) for $|\\alpha|$ approaching $1$. This feature implies strong limits to the maximum frequency allowed in our model and which will be within the reach of future gravitational-waves detectors like BBO and DECIGO \\cite{Lidsey97}, for the KHz range of frequencies. In fact, for the most consistent values of the $\\alpha$-parameter, the spectrum shows a frequency as low as in the Hz region. Concluding, these results show that, for this model, the Hz-KHz frequency range comes directly from the transition between the inflationary regime and the radiation era, and addresses important issues about the limits of the GW energy spectrum. Last but not least, we would like to highlight that one of the merits of the model we have presented for the transition from the inflationary era to the radiation dominated epoch is its relative simplicity." }, "0910/0910.2652_arXiv.txt": { "abstract": "We present a Chandra X-ray observation of the very high energy $\\gamma$-ray source HESS$~$J1640-465. We identify a point source surrounded by a diffuse emission that fills the extended object previously detected by XMM Newton at the centroid of the HESS source, within the shell of the radio supernova remnant (SNR) G338.3-0.0. The morphology of the diffuse emission strongly resembles that of a pulsar wind nebula (PWN) and extends asymmetrically to the South-West of a point-source presented as a potential pulsar. The spectrum of the putative pulsar and compact nebula are well-characterized by an absorbed power-law model which, for a reasonable $N_{\\rm H}$ value of $14\\times 10^{22} \\rm cm^{-2}$, exhibit an index of 1.1 and 2.5 respectively, typical of Vela-like PWNe. We demonstrate that, given the H$~$I absorption features observed along the line of sight, the SNR and the H$~$II surrounding region are probably connected and lie between 8 kpc and 13 kpc. The resulting age of the system is between 10 and 30 kyr. For a 10 kpc distance (also consistent with the X-ray absorption) the 2-10 keV X-ray luminosities of the putative pulsar and nebula are $L_{\\rm PSR} \\sim 1.3 \\times 10^{33} d_{10 \\rm ~kpc}^{2} \\rm erg.s^{-1}$ and $L_{\\rm PWN} \\sim 3.9 \\times 10^{33} d_{10}^{2} \\rm erg.s^{-1}$ ($d_{10} = d / 10{\\rm~kpc}$). Both the flux ratio of $L_{\\rm PWN}/L_{\\rm PSR} \\sim 3.4$ and the total luminosity of this system predict a pulsar spin-down power around $\\dot{E} \\sim 4 \\times 10^{36} \\rm erg~s^{-1}$. We finally consider several reasons for the asymmetries observed in the PWN morphology and discuss the potential association with the HESS source in term of a time-dependent one-zone leptonic model. ", "introduction": "During 2004-2006, H.E.S.S. (High energy stereoscopic system) performed a survey of the inner part of the Galaxy where its excellent capability provided a breakthrough in the field of pulsar wind nebulae (PWN) study: for the first time the morphological structure of many middle-age PWNe were resolved in the $\\gamma$-ray band. A detailed spectral and morphological analysis of the archetypal example of this population, PSR$~$B1823-13 and its associated TeV nebula HESS$~$J1825-137 revealed for the first time in $\\gamma$-rays a steepening of the energy spectrum with increasing distance from the pulsar, probably due to the synchrotron radiative cooling of the Inverse-Compton (IC) $\\gamma$-ray emitting electrons during their propagation (Aharonian et al. 2006b). In this scenario, the different sizes observed in X-rays and very high energy (VHE) $\\gamma$-rays can be explained by the different cooling timescales for the radiating electron populations (de Jager et al. 2005). However, the question of the unusual large size of this PWN candidate compared to what the dynamical simulations predict still remains (de Jager et al. 2005; de Jager $\\&$ Djannati 2008) and is difficult to solve. The confusion is even aggravated because of the non-detection of the associated supernova shell, like it is the case for most of the other middle-aged TeV PWN candidates (Gallant et al. 2007). HESS$~$J1640-465 is one of the sources discovered by HESS during this Galactic survey in 2004-2006 (Aharonian et al. 2006a). This source is marginally extended with profile close to a Gaussian shape. The spectrum is well fitted by a simple power-law with an index of $\\sim$ 2.4 and a total integrated flux above 200 GeV of $2.2 \\times 10^{-11} {\\rm ~ erg ~ cm}^{-2}{\\rm ~ s}^{-1}$. As it is shown in Figure 1, the source is spatially coincident with the SNR G338.3-0.0, known from radio observations as a $8'$ diameter broken shell with a 7 Jy flux at 1GHz (Whiteoak $\\&$ Green 1996). A radio bridge extending to the north of the system is visible in the MOST data and coincident with the bright H$~$II region G338.4+0.0 (Whiteoak $\\&$ Green 1996), possibly connected to the shell. The fact that the TeV source is center filled and doesn't match the shell has favoured the hypothesis that the emission is produced by a PWN so far. This field has been successively observed by ASCA (Sugizaki et al. 2001) and the Swift X-ray telescope (Landi et al. 2006); both detected a highly absorbed source inside the shell, with compatible positions and non thermal spectra.\\\\ In 2006, a dedicated XMM-Newton observation of the field of view revealed an extended X-ray object at the position ($16^{\\rm h}40^{\\rm m}45^{\\rm s}.4$, $-46^{\\rm o}31'31''$(J2000))(Funk et al. 2007). This extended source (XMMUJ164045.4-463131) exhibits a strongly absorbed spectrum with a potential non-thermal spectrum (index $\\sim 1.7$) centered about the VHE source position. However, because of the low number of counts of these observations the spectrum was poorly constrained, and the insufficient angular resolution was not able to resolve any compact nebula, nor a pulsar which could confirm the hypothesis of a PWN emission. We report here on the first Chandra observation of this field. We present a detailed data analysis and discuss the physical implications of this first case of a complete middle-age composite system with VHE PWN emission. ", "conclusions": "This high-resolution Chandra ACIS observation of HESS$~$J1640-165 inside the shell G338.3-0.0, has allowed us to spatially resolve a PWN and its putative pulsar, and perform a spectral analysis. We found that the spectra are both highly absorbed and non-thermal, and the indices and luminosities resemble those of typical Vela-like pulsar systems. Using H$~$I absorption features, we were able to constrain the distance of the shell between 8.5 and 13 kpc, this distance range being compatible with the large $N_{\\rm H}$ value derived from X-ray spectral analysis, we concluded that it is likely associated with the shell G338.3-0.0. We finally discussed a scenario in which the $\\gamma$-rays originate from the IC scattering and X-ray from the synchrotron emission of the electron populations in the nebula, and show that the VHE source HESS$~$J1640-165 is likely a PWN, probably associated with a Vela-like pulsar of $\\dot{E} = 4 \\times 10^{36} \\rm erg.s^{-1}$. If this interpretation is true, this object is a nice example of a middle-aged composite system where the total PWN extension (traced by the TeV emission) fills almost 75 percent of the associated SNR shell area, which is quite surprising given that dynamical simulations currently predict a 25 percent. Future search for similar cases will be useful in order to confirm or not the generality of this phenomenon and study its causes in detail. In this context, the detection of the associated pulsars will be very helpful. A next step of this study could be the high resolution radio imaging of the source and the search for pulsed emission originating from this putative pulsar position." }, "0910/0910.0723.txt": { "abstract": "{The present number of Galactic Open Clusters that have high resolution abundance determinations, not only of [Fe/H], but also of other key elements, is largely insufficient to enable a clear modeling of the Galactic Disk chemical evolution. } {To increase the number of Galactic Open Clusters with high quality measurements. } {We obtained high resolution (R$\\sim$30\\,000), high quality (S/N$\\sim$50-100 per pixel), echelle spectra with the fiber spectrograph FOCES, at Calar Alto, Spain, for three red clump stars in each of five Open Clusters. We used the classical Equivalent Width analysis method to obtain accurate abundances of sixteen elements: Al, Ba, Ca, Co, Cr, Fe, La, Mg, Na, Nd, Ni, Sc, Si, Ti, V, Y. We also derived the oxygen abundance through spectral synthesis of the 6300~\\AA\\ forbidden line. } {Three of the clusters were never studied previously with high resolution spectroscopy: we found [Fe/H]=+0.03$\\pm$0.02 ($\\pm$0.10)~dex for Cr~110; [Fe/H]=+0.01$\\pm$0.05 ($\\pm$0.10)~dex for NGC~2099 (M~37) and [Fe/H]=--0.05$\\pm$0.03 ($\\pm$0.10)~dex for NGC~2420. This last finding is higher than typical recent literature estimates by 0.2--0.3~dex approximately and in better agreement with Galactic trends. For the remaining clusters, we find: [Fe/H]=+0.05$\\pm$0.02 ($\\pm$0.10)~dex for M~67 and [Fe/H]=+0.04$\\pm$0.07 ($\\pm$0.10)~dex for NGC~7789 . Accurate (to $\\sim$0.5~km~s$^{-1}$) radial velocities were measured for all targets, and we provide the first high resolution based velocity estimate for Cr~110, $<$$V_r$$>$=41.0$\\pm$3.8~km~s$^{-1}$. } {With our analysis of the new clusters Cr~110, NGC~2099 and NGC~2420, we increase the sample of clusters with high resolution based abundances by 5\\%. All our programme stars show abundance patterns which are typical of open clusters, very close to solar with few exceptions. This is true for all the iron-peak and s-process elements considered, and no significant $\\alpha$-enhancement is found. Also, no significant sign of (anti-)correlations for Na, Al, Mg and O abundances is found. If anticorrelations are present, the involved spreads must be $<$0.2~dex. We then compile high resolution data of 57 OC from the literature and we find a gradient of [Fe/H] with Galactocentric Radius of --0.06$\\pm$0.02~dex~kpc$^{-1}$, in agreement with past work and with Cepheids and B stars in the same range. A change of slope is seen outside $R_{\\rm{GC}}$=12~kpc and [$\\alpha$/Fe] shows a tendency of increasing with $R_{\\rm{GC}}$. We also confirm the absence of a significant Age-Metallicity relation, finding slopes of --2.6$\\pm$1.1~10$^{-11}$~dex~Gyr$^{-1}$ and 1.1$\\pm$5.0~10$^{-11}$~dex~Gyr$^{-1}$ for [Fe/H] and [$\\alpha$/Fe] respectively. } ", "introduction": "Open clusters (hereafter OC) are the ideal {\\em test particles} in the study of the Galactic disk, providing chemical and kinematical information in different locations and at different times. Compared to field stars, they have the obvious advantage of being coeval groups of stars, at the same distance and with a homogeneous composition. Therefore, their properties can be determined with smaller uncertainties. Several attempts have been done in the past to derive two fundamental relations using OC: the {\\em metallicity gradient} along the disk and the {\\em age-metallicity relation} (hereafter AMR) of the disk \\citep[e.g.][]{jan79,pan80,tw97,friel02,che03,sal04}, but they were hampered by the lack of large and homogeneous high quality datasets. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% Observing Logs \\begin{table*} \\begin{minipage}[t]{17.5cm} \\caption{Observing Logs and Programme Stars Information.} \\label{obslog} \\centering \\renewcommand{\\footnoterule}{} \\begin{tabular}{l l c c c c c c c c c c c } \\hline\\hline Cluster & Star & $\\alpha_{J2000}$ & $\\delta_{J2000}$ & B & V &I$_{C}$& R & K$_S$& $n_{exp}$ & $t_{exp}^{(tot)}$ & S/N($\\simeq$6000\\AA) \\\\ & & (hrs) & (deg) &(mag) &(mag) & (mag) &(mag) &(mag) & & (sec) & \\\\ \\hline Cr~110\\footnote{Star names and I$_C$ \\& R magnitudes from \\citet{daw98}; coordinates and B \\& V magnitudes from \\citet{bra03}; K$_S$ magnitudes from 2MASS} & 2108 & 06:38:52.5 & +02:01:58.4\t &14.79 &13.35 & --- & --- & 9.76 & 6\t & 16200 & 70 \\\\ & 2129 & 06:38:41.1 & +02:01:05.5\t &15.00 &13.66 & 12.17 &12.94 &10.29 & 7 \t & 18900 & 70 \\\\ & 3144 & 06:38:30.3 & +02:03:03.0\t &14.80 &13.49 & 12.04 &12.72 &10.19 & 6\t & 16195 & 65 \\\\ NGC~2099 (M~37)\\footnote{Star names from \\citet{vanz21}; coordinates from \\citet{kiss01}; B \\& V magnitudes from \\citet{kalirai01}; I$_C$ magnitudes from \\citet{nila02}; K$_S$ magnitudes from 2MASS} & 067 & 05:52:16.6 & +32:34:45.6\t &12.38 &11.12 & 9.87 & --- & 8.17 & 3\t & 3600 & 95 \\\\ & 148 & 05:52:08.1 & +32:30:33.1\t &12.36 &11.09 & --- & --- & 8.05 & 3\t & 3600 & 105 \\\\ & 508 & 05:52:33.2 & +32:27:43.5\t &12.24 &10.98 & --- & --- & 7.92 & 3\t & 3900 & 85 \\\\ NGC~2420\\footnote{Star names from \\citet{can70}; coordinates from \\citet{stet00} and \\citet{las90}; B \\& V magnitudes from \\citet{tw90}; I$_C$ \\& R magnitudes from \\citet{stet00}; K$_S$ magnitudes from 2MASS} & 041 & 07:38:06.2 & +21:36:54.7\t &13.75 &12.67 & 11.61 &12.13 &10.13 & 5\t & 9000 & 70 \\\\ & 076 & 07:38:15.5 & +21:38:01.8\t &13.65 &12.66 & 11.65 &12.14 &10.31 & 5 \t & 9000 & 75 \\\\ & 174 & 07:38:26.9 & +21:38:24.8\t &13.41 &12.40 & --- & --- & 9.98 & 5\t & 9000 & 60 \\\\ NGC~2682 (M~67)\\footnote{Star names from \\citet{fag06}; coordinates from \\citet{fan96}; B, V \\& I$_C$ magnitudes from \\citet{san04}; B magnitude for star 286 and R magnitudes from \\citet{jan84}; K$_S$ magnitudes from 2MASS} & 0141 & 08:51:22.8 & +11:48:01.7\t &11.59 &10.48 & 9.40 & 9.92 & 7.92 & 3 \t & 2700 & 85 \\\\ & 0223 & 08:51:43.9 & +11:56:42.3\t &11.68 &10.58 & 9.50 &10.02 & 8.00 & 3 \t & 2700 & 85 \\\\ & 0286 & 08:52:18.6 & +11:44:26.3\t &11.53 &10.47 & 9.43 & 9.93 & 7.92 & 3 \t & 2700 & 105 \\\\ NGC~7789\\footnote{Star names and V \\& I$_c$ magnitudes from \\citet{gim98b}; J1950 coordinates from \\citet{kun23}, precessed to J2000; B magnitudes from \\citet{moc99}; K$_S$ magnitudes from 2MASS} & 5237 & 23:56:50.6 & +56:49:20.9\t &13.92 &12.81 & 11.52 & --- & 9.89 & 5 \t & 9000 & 70 \\\\ & 7840 & 23:57:19.3 & +56:40:51.5\t &14.03 &12.82 & 11.49 & --- & 9.83 & 6\t & 9000 & 75 \\\\ & 8556 & 23:57:27.6 & +56:45:39.2\t &14.18 &12.97 & 11.65 & --- &10.03 & 3\t & 5400 & 45 \\\\ \\hline \\end{tabular} \\end{minipage} \\end{table*} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% Observing Logs In particular, the lack of a {\\em metallicity scale} extending to solar metallicity with comparable precision to the lower metallicity regimes \\citep[i.e.,][]{zin84,car97} represents the main problem from the point of view of {\\em (i)} the study of the Galactic disk; {\\em (ii)} tests of stellar evolution models for younger and more metal-rich {\\em simple stellar populations} and {\\em (iii)} the use of those stellar populations as templates for extragalactic studies of population synthesis. Of the $\\sim$1700 known OC \\citep[][and updates]{dia02}, only a subset of $\\sim$140, i.e., 8\\% of the total, possesses some metallicity determination. Most of these have been obtained through different photometric studies in the Washington \\citep[e.g.][]{gei91,gei92}, DDO \\citep[e.g.][]{cla99}, Str\\\"omgren \\citep[e.g.][]{bru99,twa03}, UBV \\citep[e.g.][]{cam85} and IR \\citep[e.g.][]{tie97} photometric systems and passbands, giving often rise to considerable differences with those obtained from spectroscopy \\citep[see][and re\\-fe\\-ren\\-ces therein]{gra00}. In a much smaller number of clusters, abundances have been derived from low-resolution spectroscopy \\citep[e.g.,][]{ricardo,war08}, with admirable attempts to obtain large and homogeneous datasets \\citep[see][]{friel93,friel02}, in spite of the non-negligible uncertainties involved in the procedure. A few research groups (see Section~\\ref{sec-disc} for more details) are presently obtaining high quality spectra and are producing more precise abundance determinations. The study of elements other than the iron-peak ones (such as $\\alpha$, s- and r-process, light elements), allows one to put more constraints on the sites of production of those elements (SNe Ia, SNe II, giants and supergiants, Wolf-Rayet stars), and therefore on their production timescales. These are fundamental ingredients for the chemical evolution modeling of the Galactic Disk \\citep{tos82,chi01,col09}. For these reasons, we obtained high resolution spectra for a sample of poorly studied old OC. We present here the detailed abundance analysis of five clusters observed during our first run at Calar Alto. Observations and data reductions are described in Section~\\ref{sec-obs}; the linelist and equivalent width measurements are detailed in Section~\\ref{sec-li} while the abundance analysis methods and results are presented in Section~\\ref{sec-abo}; abundance results are then discussed and compared with literature results in Sections~\\ref{sec-lit}, \\ref{sec-disc} and \\ref{sec-trend}; finally, we summarize our results and draw our conclusions in Section~\\ref{sec-sum}. %__________________________________________________________________ ", "conclusions": "\\label{sec-sum} We have analyzed high resolution spectra of three red clump giants in five OC, three of them lacking any previous high resolution based chemical analysis. Given the paucity of literature data, such a small sample is enough to increase the whole body of high resolution data for OC by $\\simeq$5\\%. To compare our results with the literature, we have compiled chemical abundances based on high resolution data of 57 clusters from the literature. Given the recent and fast progress in the field, this sample is $\\sim$50\\% larger than previous literature compilations \\citep[e.g.,][]{friel02,sestito,mag09}. The main results drawn by the analysis of our five clusters are: \\begin{itemize} \\item{We provide the first high resolution based abundance analysis of Cr~110 ([Fe/H]=+0.03$\\pm$0.02 ($\\pm$0.10)~dex), NGC~2099 ([Fe/H]=+0.01$\\pm$0.05 ($\\pm$0.10)~dex) and NGC~2420 ([Fe/H]=--0.05$\\pm$0.03 ($\\pm$0.10)~dex), which only had low resolution determinations and the R$\\simeq$16000 analysis by \\citet{smi87}; our new determination of the metallicity of NGC~2420 puts this cluster in much better agreement with the global Galactic trends;} \\item{The abundances found for NGC~7789 ([Fe/H]= +0.04$\\pm$0.07 ($\\pm$0.10)~dex) and M~67 ([Fe/H]= +0.05$\\pm$0.02 ($\\pm$0.10)~dex) are in good agreement with past high resolution studies;} \\item{We provide the first high resolution based radial velocity determination for Cr~110 ($$=41.0$\\pm$3.8~km~s$^{-1}$);} \\item{We found that all our abundance ratios, with few exceptions generally explained with technical details of the analysis procedure, are near-solar, as is typical for OC with similar properties; we found solar ratios also for Na, that is generally found overabundant, and for O, which is generally found underabundant;} \\item{We do not find any significant sign of anti-correlation (or correlation) among Na, Al, Mg and O, in general agreement with past and recent results, and we can say that if such correlations indeed are present in OC, they must be much less extended than in Globular Clusters, amounting to no more than 0.2--0.3~dex at most;} \\end{itemize} With our compilation of literature data, we also could examine global Galactic trends, that are extremely useful to construct chemical evolution models for the Galaxy in general and the Galactic Thin Disc in particular. For the metallicity gradient we found a slope of --0.06$\\pm$0.02~dex~kpc$^{-1}$ considering only the high resolution data within $R_{\\rm{GC}}$=12~kpc. Our compilation contains data, including our own determinations, that fill the small gap around 9$<$R$_{\\rm{GC}}<$12 and point towards a gentle and continuous decrease, rather than a two step drop such as in \\citet{tw97} or a steep slope such as in \\citet{sestito}. We do find a flattening at $R_{\\rm{GC}}>$12~kpc and a hint of an increasing [$\\alpha$/Fe] towards the outer Disk. Concerning the AMR, we do not find any strong evidence for it, and we just note some very mild trends. If an AMR is indeed present among OC, it must be very weak and we provide upper limits to its slope both in [Fe/H] and [$\\alpha$/Fe]. %______________________________________________________________" }, "0910/0910.5843_arXiv.txt": { "abstract": "We test the theoretical prediction that the straightest dust lanes in bars are found in strongly barred galaxies, or more specifically, that the degree of curvature of the dust lanes is inversely proportional to the strength of the bar. The test used archival images of barred galaxies for which a reliable non-axisymmetric torque parameter ($Q_{\\rm b}$) and the radius at which $Q_{\\rm b}$ has been measured ($r(Q_{\\rm b})$) have been published in the literature. Our results confirm the theoretical prediction but show a large spread that cannot be accounted for by measurement errors. We simulate 238 galaxies with different bar and bulge parameters in order to investigate the origin of the spread in the dust lane curvature versus $Q_{\\rm b}$ relation. From these simulations, we conclude that the spread is greatly reduced when describing the bar strength as a linear combination of the bar parameters $Q_{\\rm b}$ and the quotient of the major and minor axis of the bar, $a/b$. Thus we conclude that the dust lane curvature is predominantly determined by the parameters of the bar. ", "introduction": "Dust lanes in galactic bars have been observed in nearby galaxies for a long time. Photographic surveys from the beginning of 20$^{\\rm th}$ century already allowed observation of these features. Pease (1917) describes the prominent and nearly straight dust lanes in NGC~5383 as a `dark streak'. The same paper presents images of NGC~1068 in which a dust lane is clearly visible. The dusty nature of these features was identified at a much later stage. Sandage (1961) writes in his atlas that 'one of the major characteristics of the SBb(s) [galaxies] is the presence of two dust lanes leaving the nucleus, one on each side of the bar and extending into the spiral arms'. Dust lanes have been recognised as related to shocks in the gas flow in barred galaxies since Prendergast (1962). Such shocks are usually found in the leading edge of the bar, roughly parallel to its major axis. The location of these shocks also corresponds to the location of areas of high shear, which prevents star formation (Athanassoula 1992; see Zurita et al.~2004 for a graphic illustration in the case of NGC~1530). Simulations show that the dust lanes are nearly straight and near the center of the bar when the galaxy has no inner Lindblad resonance (ILR), which is known to be quite a rare occurrence. ILRs cause the dust lanes to be offset from the bar major axis, and to be curved with the concavity pointing to the bar major axis (Athanassoula 1992). Simulations by Patsis \\& Athanassoula (2000) show that the higher the gas sound speed, the smaller is the offset between the dust lane and the major axis of the bar. No dust lanes are expected in nuclear bars (Shlosman \\& Heller 2002). Athanassoula (1992) predicted from simulations that dust lanes would have a greater curvature in weaker bars and that dust lanes would be nearly straight in strong bars. This effect has been empirically verified using a set of images of nine barred spiral galaxies by Knapen et al.~(2002). The aim of the present Letter is to improve the statistics compared to the latter study, and to test definitively the prediction from Athanassoula (1992). Because dust lanes can be observed so easily they offer a fundamental handle on the underlying dynamics of the galaxy and the bar, yet they have hardly been studied observationally. A basic further aim of the present work is thus to explore in a statistical fashion the dependence of dust lane shape on the dynamical properties of the galaxy which hosts them. ", "conclusions": "Using 55 galaxies with published measurements of the bar strength indicator $Q_{\\rm b}$, and for which we could obtain accurate measurements of the dust lane curvature, $\\Delta\\alpha$, we confirm the theoretical prediction made by Athanassoula (1992) that the dust lane curvature anticorrelates with the bar strength. Strong bars thus host straight dust lanes, while weaker bars can host more curved dust lanes. We do find that the anticorrelation has a large spread, but from a set of 238 numerical simulations of barred galaxies we show that this spread can be greatly reduced by using an appropriate linear combination of bar parameters ($Q_{\\rm b}$ and the axis ratio $a/b$)." }, "0910/0910.1698_arXiv.txt": { "abstract": "{}{We use the IBIS/ISGRI telescope on-board \\emph{INTEGRAL} to measure the position of the centroid of the 20-200 keV emission from the Crab region.}{We find that the astrometry of the IBIS telescope is affected by the temperature of the IBIS mask during the observation. After correcting for this effect, we show that the systematic errors on the astrometry of the telescope are of the order of 0.5 arcsec. In the case of the Crab nebula and several other bright sources, the very large number of photons renders the level of statistical uncertainty in the centroid smaller or comparable to this value.}{We find that the centroid of the Crab nebula in hard X-rays (20-40 keV) is shifted by 8.0 arcsec with respect to the Crab pulsar in the direction of the X-ray centroid of the nebula. A similar shift is also found at higher energies (40-100 and 100-200 keV). We observe a trend of decreasing shift with energy, which can be explained by an increase in the pulsed fraction. To differentiate between the contribution of the pulsar and the nebula, we divide our data into an on-pulse and off-pulse sample. Surprisingly, the nebular emission (i.e., off-pulse) is located significantly away from the X-ray centroid of the nebula.}{In all 3 energy bands (20-40, 40-100 and 100-200 keV), we find that the centroid of the nebula is significantly offset from the predicted position. We interpret this shift in terms of a cut-off in the electron spectrum in the outer regions of the nebula, which is probably the origin of the observed spectral break around 100 keV. From a simple spherically-symmetric model for the nebula, we estimate that the electrons in the external regions of the torus ($d\\sim0.35$ pc from the pulsar) reach a maximal energy slightly below $10^{14}$ eV.} ", "introduction": "The Crab nebula is the brightest astrophysical source in the $\\gamma$-ray sky ($E>30$ keV). It is the prototypical pulsar-wind nebula (PWN, see e.g., \\citet{gaensler} for a review), where material ejected from the central pulsar at different epochs interacts, producing strong shocks that accelerate electrons up to energies $\\sim10^{15}$ eV \\citep{kennel}. The synchrotron emission produced by the large population of non-thermal electrons in a magnetic field $B\\sim0.3$ mG \\citep{marsden} is detected over more than 10 orders of magnitude in the electromagnetic spectrum from radio to $\\gamma$-rays \\citep{atoyan,volpi,hesterreview}. In soft X-rays, the size of the nebula is $\\sim$2 arcmin \\citep{weisskopf}. It consists of a characteristic jet+torus structure, probably corresponding to relativistic outflows along the rotation axis and the equator of the pulsar. High-resolution \\emph{Chandra} observations also revealed the presence of an inner ring, probably associated with the conversion of the relativistic pulsar wind into a synchrotron-emitting plasma. The measured photon index decreases with radius from $\\Gamma\\sim1.6$ around the pulsar down to $\\Gamma\\sim3.0$ in the outer parts of the X-ray image \\citep{mori,willingale}, which implies that the most energetic electrons are constantly injected by the pulsar in the inner regions of the nebula. The steeper photon index in the outer regions can be explained by synchrotron cooling. This interpretation is supported by the observation of rapid X-ray variability in the inner ring \\citep{hester}, which is probably due to the presence of strong quasi-stationary shocks. The pulsar itself exhibits a hard spectrum $\\Gamma\\sim1.6$ and a period $P\\sim30$ msec with a double-peak profile. In the hard X-ray/soft $\\gamma$-ray range, early HEAO A-4 observations inferred a photon index of $\\Gamma=2.15$ \\citep{jung}. Around 100 keV a spectral break was detected, and above this energy a photon index of $\\sim$2.5 was found. This measurement agrees with the steeper spectral indices measured at higher energies by COMPTEL \\citep{strong} and EGRET \\citep{nolan}. Therefore, observations of the Crab complex around the break energy by modern experiments are important to constrain particle acceleration models. Recently, a polarized $\\gamma$-ray signal from the Crab was measured by the SPI \\citep{dean} and IBIS \\citep{forot} instruments on \\intgr. The polarization was found to be co-aligned with the spin orientation, thus indicating a possible association of the polarized signal with the inner jet. The broad-band coverage of \\intgr\\ also permitted a study of the evolution of the pulse profiles as a function of energy \\citep{mineo}. A phase-revolved analysis detected a significantly harder signal ($\\Gamma\\sim1.8$) during the interpulse phase compared to the peaks ($\\Gamma\\sim2.2$). However, the results presented there only considered the emission from the pulsar, and no information was given about the spectrum of the nebula. In this paper, we use the coded-aperture IBIS telescope \\citep{Ubeetal03} on-board \\intgr\\ \\citep{Winetal03} to measure the position of the hard X-ray/soft $\\gamma$-ray centroid of the Crab pulsar/nebula complex. In Sect. \\ref{astrom}, we analyse the astrometry of the IBIS telescope and show that the point-source localization accuracy of the instrument depends on the temperature of the mask. We correct for this effect and assess the level of systematic uncertainties in the astrometry of IBIS. In Sect. \\ref{results}, we report precise measurements of the position of the Crab pulsar/nebula and present our results. The implications of these findings are discussed in Sect. \\ref{disc}. ", "conclusions": "We have used the IBIS/ISGRI instrument on-board \\intgr\\ to measure the centroid of the Crab pulsar/nebula complex in hard X-rays/soft $\\gamma$-rays with unprecedented accuracy, with the aim of investigating the behaviour of the relativistic electron population around the break energy ($E_{\\mbox{\\tiny{break}}}\\sim100$ keV). Based on our understanding of the dependence of the astrometry of IBIS on the temperature of the mask, we have shown in Sect. \\ref{syst} that despite its moderate angular resolution (12 arcmin FWHM), in the case of sources detected with high $S/N$ IBIS/ISGRI can measure the position of astrophysical sources with an accuracy $\\sim0.5$ arcsec. Applying this method to the Crab, we found that the emission is offset significantly with respect to the position of the pulsar, by $\\sim8\"$ in the 20-40 keV band. We were also able to measure the position of the on- and off-pulse phases independently. All the measured positions are found along the line connecting the pulsar to the centroid of the X-ray emission in the torus, which corresponds roughly to the jet line (see Sect. \\ref{interp}). Performing phase-resolved imaging (see Sect. \\ref{relapos}), we found that the on-pulse emission (where the Crab pulsar accounts for an important fraction of the flux) depends on energy: the on-pulse shift with respect to the pulsar position decreases by $\\sim30\\%$ between 20 and 200 keV. This effect can be easily explained by an increase in the pulsed fraction (see Sect. \\ref{pulsfrac}). More interestingly, we find that during the off-pulse phase (where the emission from the pulsar is negligible) the centroid of the source is significantly offset from the position predicted from \\cra\\ data (see Sect. \\ref{disc}). This indicates that the morphology of the source changes around the break energy. Consistently, we show that a cut-off in the electron spectrum in the outer regions of the X-ray nebula, caused by synchrotron cooling and in agreement with the predictions of \\citet{atoyan}, can reproduce the observed shift. The centroid of the nebular emission around the break energy coincides with the inner ring, which is interpreted as the location to be the wind termination shock \\citep{weisskopf}. Therefore, our data imply that above the break energy only the strong shock regions are responsible for the emission. Comparing our results with the predictions of a simple model for the spectral evolution of the nebula, we find that the electrons in the outer regions of the torus ($d\\gtrsim0.35$ pc away from the pulsar) probably reach a maximal energy close to $10^{14}$ eV. This result agrees with theoretical studies carried out using MHD simulations \\citep{volpi}. In the near future, because of a higher angular resolution by a factor $\\sim20$ \\emph{NuSTAR} will be able to resolve the Crab pulsar/nebula complex up to $\\sim80$ keV, which will allow us to probe the electron spectrum in the different regions of the nebula close to the break energy." }, "0910/0910.4242_arXiv.txt": { "abstract": "CCD {\\it UBVRI} photometry is presented for type IIb SN 2008ax for about 320 days. The photometric behavior is typical for core-collapse SNe with low amount of hydrogen. The main photometric parameters are derived and the comparison with SNe of similar types is reported. Preliminary modeling is carried out, and the results are compared to the observed light curves. The main parameters of the hydrodynamical models are close to those used for SN IIb 1993J. ", "introduction": " ", "conclusions": "" }, "0910/0910.4418_arXiv.txt": { "abstract": "Approximately 30\\% of luminous red giants exhibit a Long Secondary Period (LSP) of variation in their light curves, in addition to a shorter primary period of oscillation. The cause of the LSP has so far defied explanation: leading possibilities are binarity and a nonradial mode of oscillation. Here, large samples of red giants in the Large Magellanic Cloud both with and without LSPs are examined for evidence of an 8 or 24 $\\mu$m mid-IR excess caused by circumstellar dust. It is found that stars with LSPs show a significant mid-IR excess compared to stars without LSPs. Furthermore, the near-IR $J$-$K$ color seems unaffected by the presence of the 24 $\\mu$m excess. These findings indicate that LSPs cause mass ejection from red giants and that the lost mass and circumstellar dust is most likely in either a clumpy or a disk-like configuration. The underlying cause of the LSP and the mass ejection remains unknown. ", "introduction": "When red giants become more luminous than $L \\sim 1000$\\,\\lsun, they begin to vary with periods of variation which fall on six physically distinct period-luminosity sequences\\footnote{On these sequences, stars on the first ascent giant branch (FGB) have slightly longer periods than those on the asymptotic giant branch (AGB) at the same luminosity \\citep{ita04,sosetal07}. Some authors designate the corresponding FGB and AGB sequences as distinct sequences.} \\citep{woo99,ita04,sosetal07,fra08}. As well as these six sequences for the most luminous red giants, there is another sequence (known as sequence-E) extending to lower luminosities and which is known to consist of close binary systems exhibiting ellipsoidal light variations \\citep{woo99,sos04}. Among the six luminous sequences all but the longest in period can be explained by radial pulsation in the fundamental and overtone modes \\citep{woo99}. The longest period sequence, sequence-D, consists of stars that exhibit variations with a short period, usually corresponding to radial pulsation in a low overtone mode, as well as a Long Secondary Period (LSP). It is the LSPs that make up sequence-D in the period-luminosity diagram for variable red giants. Approximately 30\\% of all luminous red giants have LSPs \\citep{woo99,per03,sosetal07,fra08} so the LSP phenomenon is very common in the late evolution of low mass stars. \\citet{sosetal07} suggest that LSPs can be seen at very low amplitude in up to 50\\% of luminous red giants. The LSPs also seem to occur in red supergiants \\citep{sto71}. Since the period of the LSP is approximately 4 times longer than the period of the normal-mode radial fundamental mode, the LSP can not be due to normal-mode radial pulsation. The most favoured explanations for the origin of LSPs are binarity and nonradial g-modes \\citep{woo99,hin02,wok04,der06,sos07,hin09}, but these and other explanations for the LSPs all have significant problems \\citep{wok04,nic09}. For example, stars exhibiting LSPs have velocity curves which generally are of similar shape and, in a binary model, this would indicate a favoured value for the angle of periastron of the orbit rather than the expected uniform distribution \\citep{nic09}. In addition, the full amplitudes of the velocity curves are closely concentrated around 3.5 km s$^{-1}$ \\citep{nic09} which implies the unlikely situation wherein all these stars have companions of a similar mass of $\\sim$0.09 M$_{\\odot}$. The problem with the nonradial g-mode explanation for the LSPs is that g-modes, which have the correct periods, only have substantial amplitudes in radiative regions and red giants have convective envelopes with only a thin radiative layer on top \\citep{wok04}. Currently, no satisfactory explanation for the LSPs exists. Several of the suggested explanations for the LSPs in red giants involve circumstellar dust. The change in visible light associated with the LSP is typically of a rather irregular nature, and the light variation can be large, up to a factor of two. In the binary model, it has been suggested that the light change could be due to a large dust cloud orbiting with the companion and obscuring the red giant once per orbit \\citep{woo99}. Another possible explanation for the LSPs \\citep{woo99} involves semi-periodic dust ejection events such as those predicted by theoretical models of AGB stars \\citep{win94,hof95}. Confirmation that dust was involved with the LSP phenomenon would be obtained if these stars showed a mid-infrared flux excess due to absorption of stellar light by circumstellar dust followed by re-radiation in the mid-IR. Both \\citet{hin02} and \\citet{ow03} examined the mid-IR colors of small samples of stars with LSPs in the solar vicinity using IRAS data. They found no evidence for a mid-IR excess that would indicate the presence of unusual amounts of circumstellar dust. With the completion of the Spitzer Space Telescope SAGE survey in the LMC \\citep{meix06,blum06}, it is now possible to search for the presence of a mid-IR excess in large samples of stars with LSPs, good contemporary light curves and a known distance. We now describe such a search. ", "conclusions": "We have shown that luminous red giant stars that exhibit Long Secondary Periods have larger mid-IR fluxes than similar stars without LSPs. This suggests that the LSP induces additional mass loss from the red giant, with consequent dust formation and an increase in the mid-IR flux. The mid-IR flux excess is only weakly dependent on the amplitude of the light variation of the LSP. A comparison of the near-IR $J$-$K$ color with the mid-IR $K$-$[24]$ and $K$-$[8]$ colors indicates that the dust is not in a spherically symmetric distribution, and is perhaps in a disk. Although dust is associated with the LSP phenomenon, it is still not clear what causes the LSP in the first place. In a binary model, where the velocity curves indicate predominantly eccentric orbits \\citep{nic09}, mass transfer from the red giant near periastron could lead to a circumbinary disk. A non-radial pulsation mode of low degree could also produce non-spherical mass loss and a non-spherical dust distribution. The semi-periodic dust ejection events predicted by \\citet{win94} and \\citet{hof95} produce, and indeed require, circumstellar dust. However, this phenomenon has not been found to occur in theoretical models at the low luminosities where many of the LSPs are observed (even below the tip of the FGB). Furthermore, it is also not clear why the photospheric velocity should vary with the LSP for this purely circumstellar process. One additional characteristic of LSPVs is that they have a chromosphere which varies with the LSP \\citep{wok04}. It is likely that both the chromosphere and the excess circumstellar dust are manifestations of the influence of the LSP phenomenon on matter above the stellar photosphere. Magnetic effects are a possible source of the chromosphere, but a search for magnetic fields in two solar-vicinity LSPVs has put an upper limit of 100 Gauss on magnetic fields covering more than about 10\\% of the surface of these stars \\citep{wmwn09}. Currently there is no evidence that magnetic fields are the source of the chromosphere. In summary, we have shown here that the LSP phenomenon produces excess circumstellar dust compared to stars without LSPs. Unfortunately, since all the postulated models for the LSP are potentially capable of causing mass ejection, the present detection of excess dust around LSPVs does not help us distinguish between the various models." }, "0910/0910.3530_arXiv.txt": { "abstract": "Based on a {new} estimation of their thickness, the global properties of relativistic slim accretion disks are investigated in this work. The resulting emergent spectra are calculated using the relativistic ray-tracing method, in which we neglect the self-irradiation of the accretion disk. The angular dependence of the disk luminosity, the effects of the heat advection and the disk thickness on the estimation of the black hole spin are discussed. Compare to the previous works, our improvements are that we use the self-consistent disk equations and we consider the disk self-shadowing effect. We find that at the moderate accretion rate, the radiation trapped in the outer region of the accretion disks will escape in the inner region of the accretion disk and contribute to the emergent spectra. At the high accretion rate, for the large inclination and large black hole spin, both the disk thickness and the heat advection have significant influence on the emergent spectra. Consequently, these effects will influence the measurement of the black hole spin based on the spectra fitting and influence the angular dependence of the luminosity. For the disks around Kerr black holes with $a=0.98$, if the disk inclination is greater than $60^\\circ$, and their luminosity is beyond 0.2 Eddington luminosity, the spectral model which is based on the relativistic standard accretion disk is no longer applicable for the spectra fitting. { We also confirm that the effect of the self-shadowing is significantly enhanced by the light-bending, which implies that the non-relativistic treatment of the self-shadowing is inaccurate. According to our results, the observed luminosity dependence of the measured spin suggests that the disk self-shadowing significantly shapes the spectra of GRS 1915+105, which might lead to the underestimation of the black hole spin for the high luminosity states.} ", "introduction": "The standard accretion disk model \\citep[SSD,][]{1973A&A....24..337S} and its relativistic generalization \\citep{1973blho.conf..343N} are widely used in modeling the spectral energy distribution (SED) of both active galactic nuclei (AGNs) and the galactic black hole binaries. Recently, based on the relativistic SSD, the spins of the black holes in several galactic black hole binaries are measured by fittings of their SED \\citep{2005ApJS..157..335L,2006ApJ...636L.113S,2006ApJ...652..518M,2006MNRAS.373.1004M}. One major advantage of standard accretion disk model is its simplicity, i.e., it assumes that the fluid undergoes the Keplerian motion and the energy generated by the viscous heating is radiated locally. Besides, in this model, the disk terminates at the Innermost Stable {Circular} Orbit (ISCO), and the matter inside ISCO is cold. However, when the accretion rate is relatively high ($\\dot M>0.1 \\dot M_{\\rm edd}$), in the inner region of the accretion disks, several effects will make these assumptions invalid. One such effect is the heat advection. When the vertical diffusion time scale is larger than the accretion time scale, substantial amount of heat is trapped inside the accretion flow and not able to escape to the infinity \\citep{1978MNRAS.184...53B,1982ApJ...253..873B}. This will make the local flux different from that from the SSD model and affect the profile of the radial temperature. Also, the advected heat will provide an additional pressure, which makes the disk rotation no longer Keplerian. Besides, because the heat trapped inside the accretion flow, the region inside ISCO is not cold any more. The self-consistent treatment of these effects needs to model the energy and momentum balance in the disk, one widely used method is based on the radiative MHD simulation. Although the numerical simulations have the advantage of treating the physical effects self--consistently, the huge amount of computation makes it difficult to give the qualitative fitting of the observational data. One complementary approach is to solve the energy and momentum equations of the accretion flow in terms of the ordinary differential equations (ODEs). Compared to the simulations, this method can effectively take into account the dynamical effects, while maintaining its conciseness. In this work, we use the ODE form disk equations \\citep{1981AcA....31..283P,1988ApJ...332..646A,1993ApJ...412..254C,1997MNRAS.286..681P,1998ApJ...498..313G}, and explore the solutions at both the medium and high accretion rate \\citep[where the disk is call the ``slim disk\";][]{1988ApJ...332..646A}. We use the relativistic disk equations in our work \\citep{1996ApJ...471..762A,1998MNRAS.297..739B,2003MNRAS.338.1013S} to investigate the effects of the black hole spin. The physical properties of the black hole and the accretion disks are deduced mainly from their observed SED. It is well known that the structure and the SED of the slim disk differ significantly from that of SSD \\citep{1999ApJ...516..420W,1999ApJ...522..839W}. The disk SED is influenced by a handful of effects, such as the Doppler effect, the relativistic beaming, the gravitational redshift, the gravitational -bending, and the disk self-shadowing. As the accretion rate increases, the disk becomes geometrically thick, and the disk self-shadowing \\citep{2007MNRAS.378..841W} becomes significant. In this work, we employ a ray-tracing approach \\citep[e.g.][]{1997PASJ...49..159F,1998NewA....3..647C} to explore the final spectra at the different viewing angles. Due to their disk-like geometry, the radiation from the accretion disks is far from the isotropy, and the emergent luminosity is thus angular dependent \\citep[e.g.][]{2006MNRAS.372.1208H}. For a slab of optically thick material without velocity, the disk luminosity simply varies as $L\\sim \\cos\\theta$ in the flat spacetime. In fact, because of its thickness and rotation, the angular dependence of the disk luminosity is expected to be different from that of the simple slab, further analysis is thus needed. In this paper, we calculate the angular dependence of the disk luminosity under different inclinations and the black hole spin. { The spin of GRS 1915+105 is debated in the literature \\citep{2006ApJ...652..518M,2006MNRAS.373.1004M}. By fitting the disk blackbody component, \\citet{2006ApJ...652..518M} found that the spin is near extreme, while \\citet{2006MNRAS.373.1004M} found the spin is moderate, with $a \\sim 0.8$. To show how the spin determination is influenced by the heat advection and disk self-shadowing, we re-simulate the observation carried out by \\citet{2006ApJ...652..518M} with our model and confirm the luminosity dependence of the measured spin as found by \\citet{2006ApJ...652..518M}. } In this work, the black hole mass in our canonical model is taken to be $10 M_{\\odot}$. However, due to the physical similarities between the accretion disks around the stellar-mass black holes and those around the supermassive black holes, our results can be easily generalized to the supermassive black holes and predict the SED of AGNs under the simple scaling. This paper organized as follows. In section 2, we describe briefly our disk model and the ray--tracing method. In Section\\ref{secst}-Section \\ref{secend}, we present our numerical results of the global disk structure. In Section \\ref{respec}, we present our results on the SED of the disk, focusing on the differences between the SED of the dynamical disk model and that of the SSD. The angular dependence of the total luminosity is given in Section \\ref{angulardep}. In Section \\ref{reimp}, implications for estimating the black hole spin are discussed. In Section \\ref{disc}, various observational implications of the disk self-shadowing as well as the spin of GRS 1915+105 are discussed. Finally, in Section \\ref{recon} we summarize our conclusions. ", "conclusions": "\\label{recon} In this work, we study the structure of the relativistic slim accretion disks and their emergent spectra. Besides the effect of the photon trapping, we find that when the accretion rate is appropriate, the energy trapped in the outer regions of the accretion disks can be re-radiated in the inner regions of the accretion disks. Another main result of this work is that we check the validity of the relativistic SSD. We find that at low accretion rate, such as $\\dot M \\sim 1/4$ for $a=0$ and $\\dot M \\sim 1/16$ for $a=0.98$, the relativistic SSD is still valid to describe the disk structure, and the effect of heat advection will not significantly influence the global structure of the disk. At relatively high accretion rate such as $\\dot M = 2 \\dot M_{\\rm edd}$, these effects will significantly distort the emergent spectrum, and hence affect the measurement of the black hole spin. We also find that when the accretion rate is high, the effect of the disk self-shadowing plays an even more important role than that of the heat advection alone. The emergent luminosity from the accretion disks is angular dependent, and the degree of anisotropy is dependent on the accretion rate. In this work, we investigate the angular dependence of the observed luminosity. Due to the physical similarities between the accretion disks of different black hole mass, our result can be applied to estimate the radiation anisotropy of AGNs. To check in principle the effect differences of spectra on estimation of spin, we create ``fake'' data sets with the simulated data, and fit them with the KERRBB model using XSPEC. At the relatively low accretion rate, our fitting results suggest that the simple Keplerian model is reliable. At the relatively high accretion rate, our fitting results suggest that self-shadowing effect will lead to significant under-estimation of the black hole spin. {By making our calculation relevant to the case of GRS 1915+105, without introducing additional parameters, we are able to reproduce the decrease of measured spin with the increasing luminosity, which is found in \\citet{2006ApJ...652..518M}. This finding suggests that disk self-shadowing significantly shapes the spectra of GRS 1915+105.} { There are several effects which we do not include in our calculations. It was claimed that the disk may have no time to relax to the thermodynamic equilibrium \\citep{2002ApJ...574..315O}, this effect will make our estimation of emergent flux (Equation (\\ref{rad})) invalid and make the disk thickness estimation inaccurate. In our calculations of the disk SED, we use a simplified model of the local spectrum and neglect the effect of limb--darkening and returning irradiation. Assuming a constant color correction may make our calculated spectra inadequate to be compared to the observation \\citep{2008ApJ...683..389D}, so we focus on analyzing the physical effects we have introduced. The limb-darkening effect will influence the spectra and the angular dependence of the emergent luminosity, while the returning irradiation will have some effects on both the disk structure and the emergent spectra. As \\citet{2005ApJS..157..335L} have calculated, these two effects will contribute several percents to the final results. Despite of these shortcomings, our results and the trend of the spectral evolution are still robust when the inclination angle is not too large. }" }, "0910/0910.3689_arXiv.txt": { "abstract": "At present none of galactic chemical evolution (GCE) models provides a self-consistent description of observed trends for all iron-peak elements with metallicity simultaneously. The question is whether the discrepancy is due to deficiencies of GCE models, such as stellar yields, or due to erroneous spectroscopically-determined abundances of these elements in metal-poor stars. The present work aims at a critical reevaluation of the abundance trends for several odd and even-$Z$ Fe-peak elements, which are important for understanding explosive nucleosynthesis in supernovae. ", "introduction": " ", "conclusions": "" }, "0910/0910.4583_arXiv.txt": { "abstract": "The \\emph{Fermi Gamma-Ray Space Telescope} reveals a diffuse inverse Compton signal in the inner Galaxy with a similar spatial morphology to the microwave haze observed by WMAP, supporting the synchrotron interpretation of the microwave signal. Using spatial templates, we regress out $\\pi^0$ gammas, as well as IC and bremsstrahlung components associated with known soft-synchrotron counterparts. We find a significant gamma-ray excess towards the Galactic center with a spectrum that is significantly harder than other sky components and is most consistent with IC from a hard population of electrons. The morphology and spectrum are consistent with it being the IC counterpart to the electrons which generate the microwave haze seen at WMAP frequencies. In addition, the implied electron spectrum is hard; electrons accelerated in supernova shocks in the disk which then diffuse a few kpc to the haze region would have a softer spectrum. We describe the full sky \\emph{Fermi} maps used in this analysis and make them available for download. ", "introduction": "The most detailed and sensitive maps of diffuse microwave emission in our Galaxy have been produced by the \\emph{Wilkinson Microwave Anisotropy Probe} (WMAP). An analysis of the different emission mechanisms in these maps uncovered a microwave ``haze'' towards the Galactic center (GC) that has roughly spherical morphology and radius $\\sim4$ kpc \\citep{Finkbeiner:2003im}. Since its discovery, \\cite{Finkbeiner:2004us} and \\cite{Dobler:2008ww} have argued that the microwave haze is hard synchrotron emission due to the fact that alternative hypotheses such as free-free (thermal bremsstrahlung of the ionized gas), spinning dust, or thermal dust have difficulty explaining its morphology, spectrum, or both. However, the 23-33 GHz spectrum of the haze synchrotron is harder than that expected from diffused electrons originally accelerated by supernova shocks in the plane \\citep{Dobler:2008ww}. Some variation is expected in the synchrotron spectrum, but generally in the sense that it should be \\emph{softer} at higher frequencies since the electrons lose energy preferentially at high energies as they diffuse from their source. While there are significant uncertainties in the spectrum of the haze \\citep[see the discussion in][]{Dobler:2008ww}, the data require that the diffused spectrum be roughly as hard as the expected \\emph{injection} spectrum from first-order Fermi acceleration at supernova shock fronts (number density $dN/dE \\propto E^{-2}$). That is, if the electrons were produced in shocks in the disk, then they would have to have undergone no diffusive energy losses over a $\\sim4\\pi/3 \\ (4$ kpc$)^3$ volume, which seems unlikely. Furthermore, there is significant emission in WMAP 23, 33, and 41 GHz bands from electrons that \\emph{were} generated in SN shocks; this emission has a very disk-like morphology (and softer spectral index which is consistent with shock acceleration), while the haze has a more spherical morphology.\\footnote{Hereafter ``spherical'' is taken to mean ``not disk-like'' -- if anything, the haze is non-spherical in the direction perpendicular to the disk (see \\refsec{residualmaps}).} Together, the haze spectrum and morphology imply either (1) a new class of objects distributed in the Galactic bulge and largely missing from the disk; (2) significant acceleration from shocks several kpc off the plane towards the Galactic center; or, perhaps most intriguingly, (3) a new electron component from a new physical mechanism. The claim that novel physics or astrophysics is required to explain the WMAP data is called the \\emph{haze hypothesis} to distinguish it from the two null hypotheses: (1) that the microwave haze is not synchrotron, but rather some combination of free-free and spinning dust; and (2) that the haze is synchrotron, but the electron spectrum required is not unusual. In this work we do not address the origin of the electrons, but instead consider what the data from the \\emph{Fermi Gamma-ray Space Telescope}\\footnote{See \\texttt{http://fermi.gsfc.nasa.gov/ssc/data/}} imply for their existence. \\bpm \\includegraphics[width=0.49\\textwidth]{plots/skymap-002.0-005.0GeV.eps} \\includegraphics[width=0.49\\textwidth]{plots/skymap-005.0-010.0GeV.eps} \\includegraphics[width=0.49\\textwidth]{plots/skymap-010.0-020.0GeV.eps} \\includegraphics[width=0.49\\textwidth]{plots/skymap-020.0-050.0GeV.eps} \\includegraphics[width=0.49\\textwidth]{plots/skymap-050.0-100.0GeV.eps} \\includegraphics[width=0.49\\textwidth]{plots/expmap-002.0-005.0GeV.eps} \\caption{ \\emph{Top and bottom left:} Full sky \\emph{Fermi} $\\gamma$-ray maps in various energy bins. The mask includes the 3-month \\emph{Fermi} point source catalog as well as the LMC, SMC, Orion-Barnard's Loop, and Cen A. \\emph{Bottom right:} Exposure time map with our mask overlaid and stretched to $0-100\\%$ of the peak exposure time. The variation in the exposure is a small modulation -- even setting it to unity does not change our qualitative results. } \\label{fig:fermimaps} \\epm Electron cosmic rays at $10-100$ GeV primarily lose energy in the diffuse interstellar medium by producing synchrotron microwaves and inverse-Compton (IC) scattered gammas. Synchrotron losses are proportional to magnetic field energy density, $U_B = B^2/8\\pi$, while IC losses are proportional to the interstellar radiation field energy density, $U_{ISRF}$, in the Thomson limit, and less in the Klein-Nishina limit. Bremsstrahlung off the ambient gas also occurs, but is expected to be sub-dominant in the regions of interest. Therefore, the best test of the haze hypothesis is to search for IC gammas in the \\emph{Fermi} data, which was studied in the context of dark matter signals by \\cite{Cholis:2008wq,Zhang:2008tb,Borriello:2009fa,Cirelli:2009vg,Regis:2009md,Belikov:2009cx,Meade:2009iu}. Previous studies of high latitude gamma-ray emission have reported measurements of an excess of emission above the Galactic plane, most notably SAS-2 \\citep{Fichtel:1975fi,Kniffen:1981kn}, COS-B \\citep{Strong:1984cb}, and EGRET \\citep{Smialkowski:1997sm,Dixon:1998di}. However, in each case the experiments either covered an insufficient energy range (SAS-2 and COS-B) or did not have sufficient sensitivity and angular resolution (EGRET) to permit a spatial correlation with the WMAP haze. \\emph{Fermi} overcomes both of these obstacles and allows us to search for the gamma-ray counterpart to the microwave haze for the first time. The presence of an IC signal at the expected level can confirm that the microwave haze is indeed synchrotron, ruling out the first null hypothesis. From the spectrum of the IC, we can estimate the electron spectrum required to make the signal, addressing the second null hypothesis. For example, the presence of $\\sim 50~\\GeV$ IC photons requires electrons of $E>50~\\GeV$, perhaps much greater. Furthermore, IC photons provide valuable information about the spatial distribution (disk vs.\\ bulge) of the source of these particles, which in turn can constrain hypotheses about their origin.\\footnote{The synchrotron haze depends on the Galactic magnetic field while the IC haze depends on the Galactic ISRF, and so the two morphologies should not be identical, and could be quite different.} In \\S 2 we briefly review the \\emph{Fermi} data, describe our map-making procedure, and display full-sky maps at various energy ranges. In \\S 3 these maps are analyzed with template correlation techniques and resultant residual maps and spectra are shown. Finally, \\S 4 presents our interpretation of the signals and discusses potential sources of contamination. Estimates for the possible contamination from unresolved point sources are given in \\refapp{pointsources}. Appendix \\ref{sec:datarelease} details the creation and processing of the gamma-ray sky maps used in this analysis, and provides instructions for downloading them. Appendix \\ref{sec:likelihood} contains a discussion of Poisson likelihood analysis on smoothed maps. ", "conclusions": "We have presented full sky maps generated from photon events in the first year data release of the \\emph{Fermi Gamma-Ray Space Telescope} (see Appendix \\ref{sec:datarelease} for data processing details). Using a template fitting technique, we have approximated both the spectrum and morphology of the well known gamma-ray emission components at \\emph{Fermi} energies. The SFD dust map was used to trace the $\\pi^0$ decay gammas generated by collisions of cosmic ray protons with the ISM, while the Haslam 408 MHz map was used to trace inverse Compton (IC) scattered photons from interactions of supernova shock-accelerated electrons ($\\sim 1-10$ GeV) with the interstellar radiation field (ISRF). Bremsstrahlung radiation, generated by interactions of these electrons with the ISM, should be approximately traced by some combination of these two maps. Although our template fitting technique is subject to significant uncertainties due to uncertain line of sight gas and CR distributions, a robust positive residual has been identified. This excess diffuse emission is centered on the Galactic center, and can be parameterized by a simple two-dimensional Gaussian template $(\\sigma_\\ell=15\\degree, \\sigma_b=25\\degree)$. The template-correlated spectrum of this emission is significantly \\emph{harder} than either $\\pi^0$ emission or IC from softer electrons, whose fitted spectra agree well with models. This harder spectrum coupled with the distinct spatial morphology of the gamma-ray and microwave haze are evidence that these electrons originate from a \\emph{separate component} than the softer SN shock-accelerated electrons. The gamma-ray excess is almost certainly the IC counterpart of the microwave haze excess described by \\cite{Finkbeiner:2003im} and \\cite{Dobler:2008ww}. Although it is still possible that a significant fraction is prompt photons from WIMP annihilations \\cite[e.g. the 200 GeV wino advocated by][]{Grajek:2008pg} such explanations are difficult to reconcile with the spatial similarity to the WMAP haze (see \\reffig{fermi_comp_wmap}). The simplest hypothesis is that the signal is mainly IC from the same electrons that produce the WMAP haze synchrotron. This addresses the stubborn question about the origin of the WMAP haze. Until recently, it has been argued that the WMAP haze had alternative explanations, such as free--free emission from hot gas or spinning dipole emission from rapidly rotating dust grains. However, the existence of this IC signal proves that the microwave haze is indeed synchrotron emission from a hard electron spectrum. \\emph{Fermi} LAT photon data are contaminated by particle events, especially at high energies. We have taken care to account for the isotropic background resulting from extragalactic sources, cosmic ray contamination, and heavy nuclei contamination and found that this background, though significant, is below the observed IC excess even up to 100 GeV. \\emph{Particle contamination is extremely unlikely to mimic the observed signal.} The LAT collaboration continues to refine the cuts used to define ``diffuse class'' events, and plans to release a cleaner class of events in coming months. This, along with a new public version of GALPROP, including updated ISRF models, will allow a more sophisticated analysis than that presented in this paper. We eagerly await the release of these software and data products. The spectrum and morphology of both the microwave and gamma-ray haze constrain explanations for the source of these electrons. There have been speculations that the microwave haze could indicate new physics, such as the decay or annihilation of dark matter, or new astrophysics, such as a GRB explosion, an AGN jet, or a spheroidal population of pulsars emitting hard electrons. We do not speculate in this paper on the origin of the haze electrons, other than to make the general observation that the roughly spherical morphology of the haze makes it difficult to explain with any population of disk objects, such as pulsars. The search for new physics -- or an improved understanding of conventional astrophysics -- will be the topic of future work. \\vskip 0.15in {\\bf \\noindent Acknowledgments:} We acknowledge helpful conversations with Elliott Bloom, Jean-Marc Casandjian, Carlos Frenk, Isabel Grenier, Igor Moskalenko, Simona Murgia, Troy Porter, Andy Strong, Kent Wood. This work was partially supported by the Director, Office of Science, of the U.S. Department of Energy under Contract No. DE-AC02-05CH11231. NW is supported by NSF CAREER grant PHY-0449818, and IC and NW are supported by DOE OJI grant \\# DE-FG02-06ER41417. IC is supported by the Mark Leslie Graduate Assistantship. DF and TS are partially supported by NASA grant NNX10AD85G. TS is partially supported by a Sir Keith Murdoch Fellowship from the American Australian Association. This research made use of the NASA Astrophysics Data System (ADS) and the IDL Astronomy User's Library at Goddard.\\footnote{Available at \\texttt{http://idlastro.gsfc.nasa.gov}} \\clearpage" }, "0910/0910.1190_arXiv.txt": { "abstract": "The amount of available telescope time is one of the most important requirements when selecting astronomical sites, as it affects the performance of ground-based telescopes. We present a quantitative survey of clouds coverage at La Palma and Mt.Graham using both ground- and satellite-based data. The aim of this work is to derive clear nights for the satellite infrared channels and to verify the results using ground-based observations. At La Palma we found a mean percentage of clear nights of $62.6$ per cent from ground-based data and $71.9$ per cent from satellite-based data. Taking into account the fraction of common nights we found a concordance of $80.7$ per cent of clear nights for ground- and satellite-based data. At Mt.Graham, we found a $97$ per cent agreement between Columbine heliograph and night-time observing log. From Columbine heliograph and the {\\it Total Ozone Mapping Spectrometer-Ozone Monitoring Instrument (TOMS-OMI)} satellite, we found that about $45$ per cent of nights were clear, while satellite data (GOES, TOMS) are much more dispersed than those of La Palma. Setting a statistical threshold, we retried a comparable seasonal trend between heliograph and satellite. ", "introduction": "The identification and characterization of a site for the future European Large Telescope (E-ELT) is a key issue. Moreover a quantitative survey of cloud cover for the areas selected as candidate sites for the telescope is and will continue to be an essential part of the process of site selection for future large telescopes in the same class as the E-ELT. In fact, the performance of large telescopes at optical and infra-red wavelengths is critically dependent on atmospheric cloud cover. Cloud cover is a key parameter at the time of site selection and also affects scientific output during the life of the telescope. For instance, a night-time seasonal trend of fewer clear sky can reduce regular access to the sky. Typically it is possible to quantify the presence of clouds at telescope sites using ground-based observations that provide a real time knowledge of the atmospheric condition. The fraction of clear sky can be determined using either instruments (i.e all-sky cameras) or observer estimates. Long-term records from many ground-based telescopes, which list the number of nights available for observing, are now accessible and it is possible to begin a reliable statistical study. However, this technique alone is not suitable for identifying future candidate sites where there are no telescopes. An easy evolution of this analysis is the use of meteorological satellites that provide measurements of cloud cover and other critical parameters for site testing covering large areas with different spatial and temporal resolution. Taking into account that, most of the meteorological satellites are equipped with similar instrumentation, it is not difficult to compare distinct sites observed by two or more different satellites. Additionally, since satellite data archives now cover long time periods, it is possible to have for each site the trend of these parameters in both long and short time scale. Erasmus and Sarazin (\\cite {erasmus02}) were among the first to demonstrate the successful application of satellite data for monitoring, comparison and forecasting evaluations. Erasmus and van Rooyen (\\cite {erasmus06}) quantified cloud cover at La Palma using Meteosat satellite and validating them using the ground based measurements taken at Carlsberg Meridian Telescope (CMT). In this paper we present the results of a study of cloud cover using satellite- and ground-based data obtained in two different important astronomical sites: the Observatorio del Roque de Los Muchachos (ORM) located in La Palma (in Canary Islands), hosting several international telescopes and among of them the Galileo National Telescope (TNG), and Mount Graham (in Arizona), hosting the Large Binocular Telescope (LBT). The results are compared with Erasmus and van Rooyen (\\cite {erasmus02}) and Erasmus and Sarazin (\\cite {erasmus06}). The present paper is organized as follows. In Section 2, we describe both the ground and satellite data bases adopted. In Section 3, we describe the satellite data acquisition procedure, and in Section 4 we show the data reduction procedure. Section 5 gives a discussion of the results. ", "conclusions": " \\begin{itemize} \\item \\textbf It is possible to define a threshold in satellite emissivity to select clear nights with an uncertainty of $20$ per cent. \\item \\textbf A good correlation exists between GOES 12 satellite and ground-based data. \\item \\textbf Using the common selected nights we found that $80.7$ per cent are classified clear from both GOES 12 satellite and ground logbook. \\item \\textbf The marginal increase to $81.1$ per cent of the concordance obtained including calima events confirm that the satellite is able to distinguish only the presence of clouds. \\end{itemize} A further analysis of the satellite data (e.g.a wider field, the simultaneous use of different channels, etc.) is suggest to improve the prediction of clear nights. Furthermore, we are studying the fraction of clear nights lost for high humidity or strong wind. \\subsection" }, "0910/0910.1159_arXiv.txt": { "abstract": " ", "introduction": "One of the complications in VLBI, over connected array interferometers, arises from the completely unrelated atmospheric conditions that the wavefronts propagate through before reaching the widely separated antennae. Self calibration procedures, which are the standard VLBI analysis technique for imaging radio sources, rely on closure relations to remove the station dependent complex gain factors that characterize the phase errors at each antenna. A direct detection of the source with good signal to noise ratio is required within every segment of the coherence integration time-interval. This time interval is set by the stability of the instrument and, dominantly, the atmospheric turbulence. An important consequence of the use of phase closure is that information on the absolute position of the source is lost, preventing the measurement of astrometric quantities.\\\\ The application of phase-referencing techniques, to the analysis of interleaving observations of the program source and a nearby calibrator, preserves the information on the angular separation on the sky and provides high precision relative-astrometry (Alef 1988, Beasley \\& Conway, 1995). At the observations, the scans on the scientifically interesting source, the target, are interleaved (within the coherence integration time) with observations of a calibrator source, the reference. The antenna-based corrections derived from the self-calibration analysis of the reference source observations are transferred for the calibration of the target source. Next, the target dataset is Fourier transformed without any further calibration to yield a phase referenced map of the target source, where the position offset of the peak from the center provides a precise measurement of the relative separation between both sources. The propagation of astrometric errors in the phase referencing analysis is strongly dependent on the angular separation between the target and reference sources, and range between the micro-arcsec and tenths of milli-arcsec accuracy. This phase referencing technique, from now on referred as ``conventional phase referencing'', is well established and has been used to provide high precision astrometric measurements of (relative) source positions in cm-VLBI observations. It would be highly desirable to extend this capability to the mm-VLBI regime, yet at the highest frequencies the observations are sensitivity limited: the instruments are less efficient, the sources are intrinsically weaker, and the phase coherence integration times are severely constrained by the rapid atmospheric phase fluctuations due to the variations (spatial and temporal) of the water vapor content in the troposphere. In particular, the coherence time is too short to allow an antenna to switch its pointing direction between pairs of sources, in all but the most exceptional cases (Porcas \\& Rioja, 2002), within that time range. The lack of suitable reference sources in mm-VLBI makes it almost impossible to apply ``conventional phase referencing'' techniques in the high frequency (i.e. significantly above 43\\,GHz) domain. \\\\ Therefore it would be hugely beneficial if the calibration could be performed at a lower, easier, frequency and used for data collected at a higher frequency. That is, to transfer the calibration terms (for phase/delay/rate VLBI observables) derived at a different frequency rather than at a different source as in ``conventional phase-referencing''. It should be noted that frequency switching can be performed much faster than source switching at the VLBA. Moreover the duty-cycle is now determined by the coherence time at the lower frequency. The low frequency phases provide `connection' but of course can not correct for variations faster than the duty cycle. This requires co-temporal dual frequency observations as provided by the next generation of VLBI antennae and arrays which are able to co-observe at different frequency bands, e.g. the Yebes 40m antenna and the Korean VLBI Network.\\\\ The feasibility of multi-frequency observations to correct the non-dispersive tropospheric phase fluctuations in the high frequency regime has been studied for some time. It relies on the fact that such fluctuations will be linearly proportional to the observing frequency, and hence it should be possible to use a scaled version of the calibration terms derived from the analysis of observations at a lower frequency (where more and stronger sources are available, with longer coherence integration times and better antenna performance), to calibrate higher frequency observations. It is a kind of phase referencing, between observations at two frequencies, that we call ``frequency phase transfer'' ({\\sc FPT}). Among the earliest references we found are ``Phase compensation experiments with the paired antennas method 2. Millimeter-wave fringe correction using centimeter-wave reference'' (Asaki, \\etal\\ 1998) with the Nobeyama millimeter array (NMA), and ``Tropospheric Phase Calibration in Millimeter Interferometry'' (Carilli \\& Holdaway, 1999) for application with the Very Large Array (VLA). In ``VLBI observations of weak sources using fast frequency switching'', Middelberg \\etal\\ (2005) applied this frequency phase transfer technique to mm-VLBI observations. They achieved a significant increase in coherence time, resulting from the compensation of the rapid tropospheric fluctuations, but failed to recover the astrometry, due to the remaining residual dispersive terms. Our proposed {\\sc Source/Frequency phase referencing} method endows this approach with astrometric capability for measuring frequency dependent source positions (``core-shifts'') by adding a strategy to estimate the ionospheric (and other) contributions. In ``Measurement of core-shifts with astrometric multi-frequency calibration'' (Rioja \\etal\\ 2005) we applied it to measure the ``core-shift'' of quasar 1038+528 A between S and X-bands (8.3/2.2 GHz), and validate the results by comparison with those from standard phase referencing techniques at cm-VLBI, where both methods are equivalent. \\\\ Here we present a demonstration of successful application of the {\\sc sfpr} method to astrometric mm-VLBI, a much more challenging frequency regime where conventional phase referencing fails. Also, the basis of the method, and details on the scheduling and data analysis are described. \\\\ This method opens a new horizon with targets and fields suitable for high precision astrometric studies with VLBI, especially at high frequencies where severe limitations imposed by the rapid fluctuations in the troposphere prevent the use of conventional phase referencing techniques. In addition this method can be applied to Space VLBI, where accurate orbit determination is a significant issue. This method results in perfect correction of frequency independent errors, such as those arising from the uncertainty in the reconstruction of the satellite orbit. The application to the space mm-VLBI mission VSOP-2 is described in detail in Rioja \\& Dodson (2009). ", "conclusions": "We have proposed and demonstrated a new method of astrometric VLBI calibration, suitable for mm-VLBI. It uses dual frequency observations to removed non-dispersive contributions. The additional step required to remove the ionospheric, and all other slowly varying dispersive terms, is done by including another source to cross calibrate with. Because the ionospheric patch size is very large at mm wavelengths one can use calibrators that lie a considerable distance from the source. We have presented the results from two pairs of sources, one only $14^\\prime$ apart and the other 10$^o$ apart. A single pair is sufficient to demonstrate the method, however the astrometric solution (the offset from expected position) contains the contribution from both sources (as happens in standard phase-referencing as well). This problem, however, fulfills the closure condition, so three or more sources can be used to form a closure triangle and separate the contributions from each individual source." }, "0910/0910.5066_arXiv.txt": { "abstract": "The form of the nuclear symmetry energy $E_s$ around saturation point density leads to a different crust-core transition point in the neutron star and affect the crust properties. We show that the knowledge about $E_s$ close to the saturation point is not sufficient, because the very low density behaviour is relevant. We also claim that crust properties are strongly influenced by the very high density behaviour of $E_s$, so in order to conclude about the form of low density part of the symmetry energy one must isolate properly the high density part. ", "introduction": "One of the most intriguing quantity in the description of nuclear matter in neutron stars is the symmetry energy $E_s$, which is defined as follows \\beq E_{nuc}(n,x)=V(n) + E_s(n)\\, \\alpha^2 + {\\cal O}(\\alpha^4) \\label{Enuc} \\eeq where $E_{nuc}(n,\\alpha)$ represents the energy of nucleonic matter per baryon as a function of baryon number density $n$ and the isospin asymmetry $\\alpha$, where $\\alpha\\!=\\!(n_n\\!-\\!n_p)/n$ and $n_n,n_p$ are the neutron and proton densities. At the saturation point density $n_0=0.16 \\fm3$ the value of symmetry energy corresponds to the $a_4$ parameter in the Bethe-Weizs\\\"acker mass formula, and takes the value $E_s(n_0) = 30\\pm 1\\MeV$. Isoscalar part of interactions is represented by the isoscalar potential $V(n)$ which is mainly responsible for the stiffness of the Equation of State (EoS). Density dependence of $E_s$ is however highly uncertain both below and above saturation point $n_0$. This dependence is one of the goals of the experimental investigations carried on the radioactive beam colliders \\cite{Baran:2004ih,Li:2008gp}. This kind of facilities allow for research of nuclear matter with large isospin asymmetry. The analysis shown in \\cite{Chen:2004si,Chen:2005ti,Chen:2007qb} put some constraints on the slope and curvature of $E_s(n)$ around $n_0$ however we are still far from the final conclusion about the global shape of the symmetry energy. The role played by the $E_s$ in the context of various neutron star observables was emphasized in \\cite{Lattimer:2000nx} and more detailed analysis was made in \\cite{Steiner:2004fi}. One of the first approach to the crust-core transition was presented in \\cite{Baym:1971ax} where the stability considerations were performed. This kind of analysis with an improved nuclear model was later used in \\cite{Pethick:1994ge}. In this work authors suggested the different critical density in different nuclear models comes from their discrepancy in the neutron matter description, e.g. in the symmetry energy form. The direct connection of $E_s$ to the crust-core transition point was shown in \\cite{Kubis:2006kb} and possible phase separation in the inner core was analysed in \\cite{Kubis:2007zz}. In this work we would like to go along this line to emphasize that the very low density behaviour of the symmetry energy is essential for the crust-core transition point. It is especially interesting taking into account the recent experimental results \\cite{Kowalski:2006ju,Natowitz:2010ti} which show the symmetry energy still takes large values at densities very much below $n_0$. This result is in contrast to common conviction coming from various theoretical approaches that states the $E_s$ goes almost linearly to zero for low densities. So, the consequences of the symmetry energy with such unusual feature seems to be worth to be seen. ", "conclusions": "" }, "0910/0910.2193_arXiv.txt": { "abstract": "We present an analysis of the luminosity distances of Type Ia Supernovae from the Sloan Digital Sky Survey-II (SDSS-II) Supernova Survey in conjunction with other intermediate redshift ($z<0.4$) cosmological measurements including redshift-space distortions from the Two-degree Field Galaxy Redshift Survey (2dFGRS), the Integrated Sachs-Wolfe (ISW) effect seen by the SDSS, and the latest Baryon Acoustic Oscillation (BAO) distance scale from both the SDSS and 2dFGRS. We have analysed the SDSS-II SN data alone using a variety of ``model-independent\" methods and find evidence for an accelerating universe at $>$97\\% level from this single dataset. We find good agreement between the supernova and BAO distance measurements, both consistent with a $\\Lambda$--dominated CDM cosmology, as demonstrated through an analysis of the distance duality relationship between the luminosity ($d_L$) and angular diameter ($d_A$) distance measures. We then use these data to estimate $w$ within this restricted redshift range ($z<0.4$). Our most stringent result comes from the combination of all our intermediate--redshift data (SDSS-II SNe, BAO, ISW and redshift--space distortions), giving $w = -0.81^{+0.16}_{-0.18}(\\mathrm{stat}){\\pm0.15}(\\mathrm{sys})$ and $\\Omega_M=0.22^{+0.09}_{-0.08}$ assuming a flat universe. This value of $w$, and associated errors, only change slightly if curvature is allowed to vary, consistent with constraints from the Cosmic Microwave Background. We also consider more limited combinations of the geometrical (SN, BAO) and dynamical (ISW, redshift-space distortions) probes. ", "introduction": "It is now widely believed that the late-time expansion of the universe is accelerating. General Relativity (GR) implies that the acceleration is driven by ``dark energy\" (DE) --- an unknown energy component in the universe with a negative effective pressure. Describing dark energy by an equation-of-state parameter of $w(z) = p/(\\rho c^2 )$ requires that $w<-1/3$. Alternatively, accelerated expansion could be an indication that GR is not the correct theory of gravity or that we have applied GR incorrectly in a cosmological context (see recent reviews of DE by \\citealt{2003RvMP...75..559P}, \\citealt{2006astro.ph..5313U}, \\citealt{2007AIPC..957...21C}, and \\citealt{2008arXiv0803.0982F}). Over the last decade, the most direct way of studying this acceleration of the expansion of the universe, and therefore DE, has been using Type Ia Supernovae (SNe), as they have been shown by many authors to be well-calibrated ``standard candles\" in the universe, i.e., their relative distances can be determined from the dependence of their peak luminosity on the shape of the light curve. This method was used to great effect by astronomers in 1998 to provide the first evidence for an accelerated universe (\\citealt{1998AJ....116.1009R, 1999ApJ...517..565P}; see \\citealt{2005ASSL..332...97F} for a review). Briefly, a Type Ia supernova occurs when a white dwarf star in a close binary system accretes enough mass from its companion to undergo a thermonuclear explosion in the core. Both \\citet{1993ApJ...413L.105P} and \\citet{1993PASP..105..787H} have shown that such explosions can serve as consistent light sources in the universe to high accuracy. This is achieved by transforming the measured light curve of the explosion into the rest frame of the supernova (so called K-corrections) and then correcting the luminosity at maximum as a function of the shape of the rest-frame light curve. Several techniques now exist for fitting SN light curves known under different acronyms ($\\Delta m_{15}$, \\citealt{1996AJ....112.2391H}; MLCS, \\citealt{1996ApJ...473...88R}; stretch, \\citealt{1997ApJ...483..565P}; CMAGIC, \\citealt{2003ApJ...590..944W}; BATM, \\citealt{2003ApJ...594....1T}; SALT, \\citealt{2005A&A...443..781G}; $\\Delta C_{12}$, \\citealt{2006ApJ...645..488W}; SALT2, \\citealt{2007A&A...466...11G}; SiFTO, \\citealt{2008ApJ...681..482C}). In this analysis, we consider MLCS2k2 \\citep{2007ApJ...659..122J, 1996ApJ...473...88R}, which is among the most commonly used, and best tested, of these various techniques. Recently, several dedicated SN surveys have been carried out to confirm and extend the earlier detections of an accelerating universe (HST, \\citealt{2004ApJ...607..665R,2007ApJ...659...98R}; SNLS, \\citealt{2006A&A...447...31A}; ESSENCE, \\citealt{2007ApJ...666..694W}) as well as new compilations of existing SN datasets \\citep{2007ApJ...666..716D, 2008arXiv0804.4142K, 2009arXiv0901.4804H}. In addition to supernovae, observations of Baryon Acoustic Oscillations (BAO) can be used to measure distances in the universe \\citep{2003ApJ...594..665B, 2003ApJ...598..720S, 2003PhRvD..68f3004H}. The BAO are caused by sound waves in the early universe which leave a preferred scale in the distribution of matter equal to the sound horizon at recombination. Today, this scale corresponds to $\\sim100/h$ Mpc (Hubble constant at present: $H_0 = 100h$~km/s/Mpc) and can thus be used as a ``standard ruler\" throughout the universe. The BAO signature has been detected in the clustering of galaxy clusters by \\citet{2001ApJ...555...68M}, in the correlation of galaxies in the Sloan Digital Sky Survey (SDSS, \\citealt{2000AJ....120.1579Y}) by \\citet{2005ApJ...633..560E}, \\citet{2006AA...459..375H}, \\citet{2007MNRAS.378..852P}, and \\citet{2007MNRAS.374.1527B}, and in the Two-degree Field Galaxy Redshift Survey (2dFGRS, \\citealt{2001MNRAS.328.1039C}) by \\citet{2005MNRAS.362..505C}. In addition to the geometrical methods discussed above, observations of the dynamical properties of matter can provide constraints on the matter density of the universe and $w$ (assuming General Relativity is the appropriate theory of gravity). For example, the growth rate of structure in the universe can be observed via the coherent infall of galaxies into large clusters and superclusters of galaxies seen in redshift surveys \\citep{1987MNRAS.227....1K}. Also, the growth of structure can be measured using the late-time Integrated Sachs-Wolfe (ISW) effect \\citep{1967ApJ...147...73S}, which has now been detected to high significance through the cross-correlation of galaxy surveys with the Cosmic Microwave Background (CMB) (see \\citealt{2008PhRvD..77l3520G} and references therein). The ISW is sensitive to deviations from a matter-dominated, Einstein-de Sitter universe ($\\Omega_M=1$, where $\\Omega_M$ is the matter density at present divided by the critical density.). Taken together, the present combination of cosmological measurements suggests we live in a flat univers, dominated by a cosmological constant ($\\Lambda$) with the energy density in matter and $\\Lambda$ known to a statistical accuracy of better than a few percent \\citep[see][]{2009ApJS..180..306D}. However, several of these cosmological measurements, especially SNe, are becoming limited by their systematic uncertainties which are now dominating, e.g., \\citet{2009arXiv0901.4804H} showed that the best combination of available SNe and BAO measurements provide $1+w=0.013^{+0.066}_{-0.068}$ but with a systematic uncertainty of $0.11$. Therefore, it is clear that future cosmological surveys must resolve these systematic errors through new observations and better analysis methods to mitigate their effect. This paper is one of three complementary papers focused on the cosmological analysis of a new sample of intermediate supernova distances recently obtained by the SDSS-II Supernova Survey (see Section \\ref{chapter_1st_year_data} for details). Our analysis differs from those presented in our companion papers of \\citet{kessler} and \\citet{davis2008}, as we first study the cosmological information obtained solely from the SDSS-II SN sample, and then in combination with other cosmological probes over the same redshift range ($z<0.4$). Alternatively, \\citet{kessler} presents a detailed examination of the impact of both statistical and systematic errors on deriving standard cosmological constraints based on the combination of the SDSS-II SN with most of the currently available high and low redshift SNe and which are all analysed in a consistent way. \\citet{davis2008} then use the same compilation of data to study an expanded set of exotic cosmological models, in combination with a wider variety of other cosmological information. Our approach is also complementary to the many other analyses in the literature (e.g. \\citealt{2007ApJ...666..716D}, \\citealt{2008arXiv0804.4142K}, \\citealt{2009arXiv0901.4804H}) that have used data from all possible sources. In our approach we concentrate on the information from cosmological measurements that cover the same range of redshifts as the SDSS SN sample. Our aim is not to derive the most stringent limit on cosmological parameters available, but rather to verify that if we restrict ourself to probes coming from a small and similar redshift slice the results on cosmological parameters remain stable and consistent. This approach is warranted because of the growing emphasis on controlling systematic uncertainties in the analysis of cosmological data. There are clearly a number of systematic uncertainties that could affect the use of SNe as cosmological probes which likely depend on, or change with, redshift, including SN evolution (e.g., changes in the metallicity of progenitor stars \\cite{2003ApJ...590L..83T, 2009ApJ...691..661H, 2009ApJ...693L..76S}), intergalactic dust \\citep{2007ApJ...664L..13C, 2008MNRAS.386..475H}, Malmquist bias and the effects of gravitational lensing and peculiar velocities \\citep{2006PhRvD..73l3526H}. Moreover, the photometric uncertainties associated with combining SN data from multiple surveys, over a range of redshifts, is already seen as the main limitation in using presently available datasets \\citep[see][]{2009arXiv0901.4804H}. Our analysis addresses this issue by focusing exclusively on the SDSS SN sample, which is derived from a well-understood and stable photometric system. The SDSS has a relative photometric accuracy of better than $2\\%$ in griz, and $4\\%$ for the u--band (Padmanabhan et al. 2007), while the absolute calibration is also known to be of the order of $1\\%$, leading to a homogeneous set of SN light-curves with high photometric accuracy \\citep[see][]{2008AJ....136.2306H}. This set of data is robust to uncertainties in light-curve fitting. For example, the MLCS2k2 and SALT2 fits to the SDSS-only sample are shown to agree well \\citep[see Section 10 in][]{kessler} which is not the case when the higher redshift SN samples are added. The outline of this paper is as follows. In Section \\ref{chapter_1st_year_data}, we describe the SDSS-II SN data and use that data in Section 3 to study the cosmic acceleration in the Universe. Section 4 then compares the SDSS-II SN luminosity distances to the BAO distances from the SDSS and 2dFGRS, checking the distance duality relation. We then derive in Section 5 constraints on $w$ by combining the best-fit luminosity distances for SDSS-II SNe with the growth rate of structure measurements taken from \\citet{2003MNRAS.346...78H} and a new measurement of the ISW effect taken from \\citet{2008PhRvD..77l3520G}. We conclude in Section \\ref{conclusions}. \\begin{figure} \\center \\includegraphics[width=80mm] {fig1.eps} \\caption{Residual Hubble diagram with respect to an empty universe for the 103 Type Ia SNe from the first year of operation of the SDSS-II SN Survey. The red line shows a $\\Lambda$CDM model with $(\\Omega_M, \\Omega_\\Lambda) = (0.3, 0.7)$, similar to our best-fit model given in Table \\ref{table_w}. The blue line is the best fit to the data derived from the ``sliding window\" technique described in Section \\ref{sec:nonpa}, and the cyan and green shaded regions correspond to the $1\\sigma$ and $2\\sigma$ confidence intervals, respectively. The black line indicates an expansion history with a deceleration parameter of $q_0 = -0.34$ as described in Section \\ref{section_global_fit}. The lower panel shows the same parameterization but now converted to $D_V$ according to Eq. (\\ref{equation_dv}). The two data points represent the BAO measurements from \\citet{2009arXiv0907.1660P}.} \\label{fig_sdss_hubble} \\end{figure} ", "conclusions": "\\label{conclusions} We present an analysis of the luminosity distances of Type Ia Supernovae from the Sloan Digital Sky Survey-II (SDSS-II) Supernova Survey in conjunction with other intermediate redshift ($z<0.4$) cosmological measurements including redshift-space distortions from the 2dFGRS, the ISW effect, and the BAO distance scale from both the SDSS and 2dFGRS. We have analyzed the SDSS-II SN luminosity distances using several 'model-independent' methods, including fitting the data using a $q(z)$ parameterization, principal components, and a non-parametric ``sliding window\" method. We find consistent results between all these methods that provides evidence for an accelerating universe based solely on the first-year SDSS-II SN data. The strongest evidence we find comes when we make the strongest assumptions, that $q_0$ is constant and the universe is flat which gives probability for acceleration of $>97\\%$. We also compare our SDSS-II SN data with the local BAO measurements, and find they are in good agreement. This is in contrast with the findings of \\citet{2007MNRAS.381.1053P} which found tension between the two distance measures, but confirms the new BAO analysis of Percival et al. (2009) who note that this tension has now lessened. Taking this observation further, we test the distance duality relation, i.e., for any metric theory of gravity, we expect $d_L/(d_A (1+z)^2) = 1$. We see no evidence for a discrepancy from this relation (at the one sigma level) in contrast to previous claims for a potential violation on the $2\\sigma$ level as seen in \\citep{2004PhRvD..69j1305B, 2008JCAP...07..012L}. Finally, we present a new measurement of the equation-of-state parameter of dark energy using a combination of geometrical distances in the universe and estimates for the growth rate of structure. Our strongest constraint comes from the combination of all four data-sets discussed herein (SDSS-II SN, BAO, redshift-space distortions, ISW) with $w=-0.81^{+0.16}_{-0.18}(stat)$ and $\\Omega_M=0.22^{+0.09}_{-0.08}(stat)$ (assuming a flat universe). However, the combination of just the SDSS-II SNe and the ISW measurements alone is almost as powerful in constraining these parameters (Table 1). Our results only change slightly if we allow curvature to vary, consistent with the CMB measurements (see Appendix \\ref{appendix_curvature}). We quote a systematic uncertainty of $\\Delta w =\\pm0.15$ based on the details of the MLCS2k2 light--curve fitter (see \\citealt{kessler} for a fuller discussion). Thus we have shown that low-redshift cosmological probes give a self-consistent picture of the distance-redshift relation. When combined with growth of structure and ISW at the same epoch that picture is consistent with $\\Lambda$CDM and re-enforces the complementarity amongst other data and analyses in the literature." }, "0910/0910.5857_arXiv.txt": { "abstract": "We present a transport model that describes the orbital diffusion of asteroids in chaotic regions of the 3-D space of proper elements. Our goal is to use a simple random-walk model to study the evolution and derive accurate age estimates for dynamically complex asteroid families. To this purpose, we first compute local diffusion coefficients, which characterize chaotic diffusion in proper eccentricity ($e_p$) and inclination ($I_p$), in a selected phase-space region. Then, a Monte-Carlo-type code is constructed and used to track the evolution of random walkers (i.e.\\ asteroids), by coupling diffusion in ($e_p$,$I_p$) with a drift in proper semi-major axis ($a_p$) induced by the Yarkovsky/YORP thermal effects. We validate our model by applying it to the family of (490) Veritas, for which we recover previous estimates of its age ($\\sim 8.3$~Myr). Moreover, we show that the spreading of chaotic family members in proper elements space is well reproduced in our randomk-walk simulations. Finally, we apply our model to the family of (3556) Lixiaohua, which is much older than Veritas and thus much more affected by thermal forces. We find the age of the Lixiaohua family to be $155\\pm 36$~Myr. ", "introduction": "\\label{} As first noted by \\citet{hirayama1918}, asteroids can form prominent groupings in the space of orbital elements. These groups, nowadays well-known as \\emph{asteroid families}, are believed to have resulted from catastrophic collisions among asteroids, which lead to the ejection of fragments into nearby heliocentric orbits, with relative velocities much lower than their orbital speeds. To date, several tens of families have been discovered across the whole asteroid Main Belt \\citep[e.g.][]{bend02,nes2006}. Also, families have been identified among the Trojans \\citep{milani93,beauge01}, and most recently, proposed to exist in the Transneptunian region \\citep{brown07}. Studies of asteroid families are very important for planetary science. Families can be used, e.g.\\ to understand the collisional history of the asteroid Main Belt \\citep{bottke05}, the outcomes of disruption events over a size range inaccessible to laboratory experiments \\citep[e.g.][]{michel03,durda07}, to understand the mineralogical structure of their parent bodies \\citep[e.g.][]{cellino02} and the effects of related dust ``showers'' on the Earth \\citep{farley06}. Obtaining the relevant information is, however, not easy. One of the main complications arises from the fact that the age of a family is, in general, unknown. Thus, accurate dating of asteroid families is an important issue in the asteroid science. A number of age-determination methods have been proposed so far. Probably the most accurate procedure, particularly suited for young families (i.e.\\ age $<10~$Myr), is to integrate the orbits of the family members backwards in time, until the orbital orientation angles cluster around some value. As such a conjunction of the orbital elements can occur only immediately after the disruption of the parent body, the time of conjunction indicates the formation time. The method was successfully applied by \\citet{nes2002,nes2003} to estimate the ages of the Karin cluster (5.8$\\pm$0.2 Myr) and of the Veritas family (8.3$\\pm$0.5 Myr). This method is however limited to groups of objects residing on regular orbits. For older families (i.e.\\ age $>100~$Myr), one can make use of the fact that asteroids slowly spread in semi-major axis due to the action of Yarkovsky thermal forces \\citep{farvok99}. As small bodies drift faster than large bodies, the distribution of family members in the $(a_p,H)$ plane -- where $a_p$ is the proper semi-major axis and $H$ is the absolute magnitude -- can be used as a clock. That method was used by \\citet{nes2005} to estimate ages of many asteroid families. In these estimations the initial sizes of the families were neglected, so that this methodology can overestimate the real age by a factor of as much as 1.5 -2. An improved version of this method, which accounts for the initial ejection velocity field and the action of YORP thermal torques, has been successfully applied to several families by \\citet{vok06a} and \\citet{vok06b}. Again, it is not straightforward to apply this method to families located in the chaotic regions of the asteroid belt. \\citet{MilFar94} suggested that asteroid families, which reside in chaotic zones, can be approximately dated by \\emph{chaotic chronology}. This method is based on the fact that the age of the family cannot be greater than the time needed for its most chaotic members to escape from the family region. In its original form, this method provides only an upper bound for the age. Recently, \\citet{menios07} introduced an improved version of this method, based on a statistical description of transport in the phase space. \\citet{menios07} successfully applied it to the family of (490) Veritas, finding an age of 8.7$\\pm$1.7~Myr, which is statistically the same as that of 8.3$\\pm$0.5 Myr, obtained by \\citet{nes2003}\\footnote{Of course, Tsiganis et al.\\ (2007) and Nevorn\\'y et al.\\ (2003) used a chaotic and a regular subsets of the family, respectively.}. Despite these improvements, the chaotic chronology still suffered from two important limitations. It did not account for the variations in diffusion in different parts of a chaotic zone, which can significantly alter the distribution of family members (i.e.\\ the shape of the family). Moreover, it did not account for Yarkovsky/YORP effects, thus being inadequate for the study of older families. In this paper we extend the chaotic chronology method, by constructing a more advanced transport model, which alleviates the above limitations. We first use the Veritas family as benchmark, since its age can be considered well-defined. Local diffusion coefficients are numerically computed, throughout the region of proper elements occupied by the family. These local coefficients characterize the efficiency of chaotic transport at different locations within the considered zone. A Monte-Carlo-type model is then constructed, in analogy to the one used by \\citet{menios07}. The novelty of the present model is that it assumes variable transport coefficients, as well as a drift in semi-major axis due to Yarkovsky/YORP effects, although the latter is ignored when studying the Veritas family. Applying our model to Veritas, we find that both (a) the shape of its chaotic component and (b) its age are correctly recovered. We then apply our model to the family of (3556) Lixiaohua, another outer-belt family but much older than Veritas and hence much more affected by the Yarkovsky/YORP thermal effects. We find the age of the Lixiaohua family to be $\\sim 155$~Myr. We note that, depending on the variability of diffusion coefficients in the considered region of proper elements, this new transport model can be computationally much more expensive than the one applied in \\citet{menios07}. This is because, if the values of the diffusion coefficients vary a lot across the considered region, one would have to calculate them in many different points. However, even so, this computation needs to be performed only once. Then, the random-walk model can be used to perform multiple runs at very low cost, e.g.\\ to test different hypotheses about the original ejection velocities field or about the physical properties of the asteroids. On the other hand, for ``smooth\" diffusion regions in which the coefficients only change by a factor of 2-3 across the considered domain, the model can be simplified. In such regions, the age of a family can be accurately determined even by assuming an average (i.e.\\ constant over the entire region) diffusion coefficient, as we show in Section 3. ", "conclusions": "\\label{} We have presented here a refined statistical model for asteroid transport, which accounts for the local structure of the phase-space, by using variable diffusion coefficients. Also, the model takes into account the long-term drift in semi-major axis of asteroids, induced by the Yarkovsky/YORP effects. This model can be applied to simulate the evolution of asteroid families, also giving rise to an advanced version of the \"chaotic chronology\" method for the determination of the age of asteroid families. We applied our model to the Veritas family, whose age is well constrained from previous works. This allowed us to assess the quality and to calibrate our model. We first analyzed the local diffusion characteristics in the region of Veritas. Our results showed that local diffusion coefficients vary by about a factor of $2-3$ across the $(e_p,\\sin I_p)$ region covered by the (5,-2,-2) MMR. Thus, although local coefficients are needed to accurately model (by the MCMC method) the evolution of the distribution of group-A members, average coefficients are enough for a reasonably accurate estimation of the family's age. We note though that the variable coefficients model reduces the error in $\\tau$ by $\\sim 30\\%$, but requires a computationally expensive procedure. Using the variable coefficients MCMC model, we found the age of the Veritas family to be $\\tau$ = (8.7$\\pm$1.2)~Myr; a result in very good agreement with that of \\citet{nes2003} and \\citet{menios07}. We used our model to estimate also the age of the Lixiaohua family. This family is similar to the Veritas family in many respects; it is a typical outer-belt family of C-type asteroids, crossed by several MMRs. Like the Veritas family, only the main chaotic zone (MCZ) shows appreciable diffusion in both eccentricity and inclination. On the other hand, this is a much older family and the Yarkovsky effect can no longer be ignored. This is evident from the distribution of family members, adjacent to the MCZ. Our model suggests that the age of this family is between 100 and 230~Myr, the best estimate being $155\\pm 36~$Myr. Note that the relative error is $\\approx 23\\%$, i.e.\\ close to the $\\approx 20\\%$ that \\citet{menios07} found for the Veritas case, using a constant coefficients MCMC model. Our model shares some similarities with the Yarkovsky/YORP chronology. Both methods are basically statistical and make use of the quasi-linear time evolution of certain statistical quantities (either the spread in $\\Delta a$ or the dispersion in $e_p$ and $\\sin I_p$), describing a family. There are, however, important differences. The Yarkovsky/YORP chronology method works better for older families and the age estimates are more accurate for this class of asteroid families (provided there are no other important effects on that time scale). On the other hand, for our method to be efficient, we need that diffusion is fast enough to cause measurable effects, but slow enough so that most of the family members are still forming a robust family structure (i.e. there is no dynamical ``sink'' that would lead to a severe depletion of the chaotic zone). Thus, our model can be applied to a limited number of families that reside in complex phase-space regions, but, in the same time, this is the only model that takes into account the chaotic dispersion of these families. There are at least a few families for which both chronology methods can be applied, thus leading to more reliable age estimates, as well as to a direct comparison of the two different chronologies. For example, the families of (20) Massalia and (778) Theobalda would be good test cases. We, however, reserve this for future work. An important advantage of the model is that it can be used to estimate the physical properties of a dynamically complex asteroid family, provided that its age is known by independent means (e.g.\\ by applying the method of \\citet{nes2003} to the regular members of the family). A large number of MCMC runs can be performed at low computational cost, thus allowing a thorough analysis of the physical parameters of family members or the properties of the original ejection velocities field that better reproduce the currently observed shape of the family." }, "0910/0910.0196_arXiv.txt": { "abstract": "The disruptive effect of galactic tides is a textbook example of gravitational dynamics. However, depending on the shape of the potential, tides can also become \\emph{fully} compressive. When that is the case, they might trigger or strengthen the formation of galactic substructures (star clusters, tidal dwarf galaxies), instead of destroying them. We perform $N$-body simulations of interacting galaxies to quantify this effect. We demonstrate that tidal compression occurs repeatedly during a galaxy merger, independently of the specific choice of parameterization. With a model tailored to the Antennae galaxies, we show that the distribution of compressive tides matches the locations and timescales of observed substructures. After extending our study to a broad range of parameters, we conclude that neither the importance of the compressive tides ($\\approx 15\\%$ of the stellar mass) nor their duration ($\\sim 10^7 \\U{yr}$) are strongly affected by changes in the progenitors' configurations and orbits. Moreover, we show that individual clumps of matter can enter compressive regions several times in the course of a simulation. We speculate that this may spawn multiple star formation episodes in some star clusters, through e.g., enhanced gas retention. ", "introduction": "That the violent collision at the origin of a galaxy merger acts to develop strong tidal features is a long-acknowledged fact of its evolution in time, as encapsulated in the Toomre sequence \\citep[see several illustrations in][]{Sandage1961, Toomre1972, Toomre1977, Hibbard1996, Laine2003, Elmegreen2007}. The burst of star formation triggered during such an event is a catalyst for the evolution of entire galactic stellar populations \\citep{Barton2000, Bridge2007, Li2008}. This process would be even more significant at high redshift ($z \\gtrsim 1$, \\citealt{Madau1996}), through either repeated accretion during the formation of a major galaxy, or enhanced merger rates driven by the formation of cosmological large-scale structures. However, the lack of resolution at these redshifts is an intrinsic difficulty when trying to identify stellar associations down to a few thousand solar masses, of a size spanning a few parsecs. At today's resolution $\\sim 0.01\\arcsec$, a full coverage of all modes of star formation triggered during a merger limits surveys to the low redshift, local Universe. Over the past decade, high-resolution HST observations of the proto-typical merger NGC 4038/39 (the Antennae) have revealed a surprisingly high number of dense, massive young clusters of stars in its central region \\citep[][]{Whitmore1995, Meurer1995, Mengel2005, Whitmore2007}. A key question which emerges from such studies pertains to the likelihood that stellar associations will survive in the rapidly-varying background potential of the merger, once they have outlasted the early phase of internal evolution \\citep{Elmegreen1997, Bastian2006, Gilbert2007}. Based on samples of some $10^3$ Antennae clusters, \\citet{Fall2005} derived a feature-less power-law age distribution ($dN/d\\tau \\propto \\tau^{-1}$) with a median value of $\\sim 10^7 \\U{yr}$, a trend they argued to be independent of mass. Their analysis may be interpreted as implying that young clusters dissolve rapidly into their gas cloud nursery, prior to reaching virial equilibrium \\citep{Whitmore2007,deGrijs2007}. \\citet{Gieles2009} has shown that mass-dependent models of cluster evolution leading to dissolution would inevitably introduce a feature in the age distribution of Antennae clusters, lending support to the conclusions of \\citet{Fall2005}. Since cluster dissolution in equilibrium galaxies invariably lead to Gaussian distributions \\citep{Vesperini1997, Baumgardt2003}, one is drawn to ask whether the power-law properties of young cluster distributions in rapidly evolving, out-of-equilibrium mergers are shaped mainly by gravitational effects, or hydrodynamics, or both. Clearly the evolution of the large-scale gravitational potential imprints the smaller scales of star formation controlled by the hydrodynamics of accretion seeds \\citep{Klessen1998, MacLow2004}. Knowing how to connect the two would be a major step toward a global understanding of the evolution of stellar populations in mergers. One way of addressing this problem consists in recreating observations numerically. Simulations of galaxy mergers present much more complicated kinematics that increase the complexity level of star formation modeling when compared, say, to galaxies taken in isolation. To resolve the many star-forming regions in a galaxy merger, several studies rely on smoothed particle hydrodynamics \\citep[SPH,][]{Gingold1977} simulations, assuming a star formation recipe mainly based on a local density or pressure threshold \\citep{Mihos1996, Bekki2002a, Bekki2002b, Springel2003, Li2004, Robertson2008, Karl2008, Johansson2009}. Other approaches like sticky particles \\citep[e.g.][]{Bournaud2008} or adaptive mesh refinement \\citep{Kim2009} are also employed to investigate the formation of substructures (star clusters or tidal dwarf galaxies, TDGs) at high resolution. \\citet{Barnes2004} compared the common SPH star formation rule with a new prescription based on the energy dissipation through shocks, with an application to NGC 4676 (the Mice). He showed that to gauge correctly the radiative energy from shocks and to recover a more realistic star-formation rate, it is necessary to take into account the divergence of the background velocity field, a strong hint that large-scale motion bears down on the small-scale physics of star formation. \\citet{Dobbs2009} recently showed that the surface density of neither the total gas (\\ion{H}{1} + H$_2$) nor the warm gas only (\\ion{H}{1}) were good tracers for the local star formation rate, considering an isolated disk. They also underlined that a strong spiral shock would increase the local density, but the velocity dispersion too, which preempts a one-to-one relation between star formation (averaged over the whole disk) and shocks. These results suggest to consider (on top of a density contrast) dynamical properties such as the convergence of the velocity field, either through hydrodynamical shock dissipation or through a background gravitational potential. On a smaller scale, several studies starting with \\citet{Ebert1955} and \\citet{Bonnor1956}, investigated the hydrodynamics of gaseous systems like giant molecular clouds confined by an external pressure \\citep[see e.g.][]{Kumai1986, Elmegreen1989, Harris1994, Elmegreen1997}. They showed that such an external effect could trigger the collapse of the cloud, so linking galactic-scales hydrodynamics and cluster-size phenomena. Despite this impressive body of work, the high-resolution needed to link gravitational galactic dynamics to the very small scales of star formation is still out-of-reach of current 3D numerical models. A possible way out of this dilemma would be to impose the right boundary conditions on a small volume within which high-resolution hydro codes may resolve the physics of star formation. This paper takes a step in that direction and investigates how these boundary conditions might evolve as a function of time. \\citet{Renaud2008} already explored the evolution of the tidal field during a merger event, with a model tailored to the Antennae galaxies. They showed that \\emph{fully} compressive tides (see Section~\\ref{sec:tidaltensor}) develop during the interaction of two disk galaxies and last for few $\\times 10^7 \\U{yr}$. They concluded that such a tidal mode could play an important role in triggering the formation of star clusters, or at least prevent them from destruction. Here, we detail further and extend these results through a larger sample of progenitor galaxies and merger orbits. It is hoped that this more general yet precise description of the tidal field will in turn allow more realistic boundary conditions for volumes where star formation takes place. In Section 2, we present analytically how compressive tides may develop when two potentials overlap, as in the case of a major galaxy merger. Section 3 describes the application to $N$-body simulations and the limitations of the method. Section 4 details the Antennae reference model and Section 5 makes direct comparisons between the parameter survey and this reference. We close with a discussion on the link between the physical process described and the observational facts. ", "conclusions": "\\subsection{Statistics of the tidal field} The gravitationnal tidal field of galaxies is often presumed to disrupt smaller bodies orbiting in their vicinity. This point of view stems from a wealth of observational data of on-going disruption of tidal dwarf galaxies, as well as strong theoretical underpinning going all the way back to Roche and the notion of an outward stretch operating on finite-size bodies. But a close inspection of the full tidal field tensor leads one to conclude that in several realistic situations, such as in the core of flat-top density profiles or the mid-point of spiral disks, the tide will work in reverse and act to shelter structures from dissolution. Because the character of the tidal field is sensitive to the presence of clumps or other sub-structures, it seems natural to seek out quantitatively the evolution of the tides in the context of a major galaxy mergers, when large structures and smallish clumps will form in great numbers. A key question that we have addressed in this paper is whether compressive tidal modes could widely influence the early life of star clusters or TDGs, both in terms of their location in the merger, and their duration in time. The reference Antennae model discussed earlier shows that fully compressive tidal modes spread throughout the merger, including in key areas such as the nuclei and the tidal tails. The statistics of the duration of these modes reveal that most of them last long enough to impact on the formation of bound substructures on a scale of $\\sim 200 \\U{pc}$ or smaller. This conclusion was confirmed in a parameter survey of galaxy mergers, which showed that the shape of the tidal mode distribution function is by far unaffected by details of the systems undergoing a merger (spin-orbit coupling, mass ratio, impact parameter, and so on). Broadly speaking, compressive tidal modes always develop to reach the same level of intensity in all the simulations that we have performed. However, some trends have been highlighted. First of all, the mass fraction entering compressive modes scales with the available mass: a low-mass progenitor undergoing a collision yields a low mass-fraction of compressive tidal modes. This is clear from the spin- and mass ratio explorations. Secondly, an encounter with a small pericenter distance brings more material to the collision in a smaller volume. As a result, the number of mass elements experiencing compressive tidal modes is significantly higher for close passages. \\subsection{Tidal field and star formation rate} For every interaction leading to mass exchange and possibly a merger, the fraction of bodies experiencing a compressive tidal field rises by a factor of at least 4, and up to 13 in the extreme case of a face-on collision, when compared with galaxies in isolation. These figures match well with the results of \\citet{diMatteo2008} who presented a study of the SFR over a large set of mergers, using both SPH and sticky particles techniques. We also remark that the starburst activity derived from their simulations would not last longer than $\\sim 500 \\Myr$, which matches well with our results. This provides indirect evidence of a close link between fluctuations in the global galactic gravitational field, and the small-scale volumes where stars are presumed to form \\citep[recall Fig.~2 of][for another, complementary argument]{Renaud2008}. \\Citet{diMatteo2008} highlighted that retrograde encounters enhance the SFR of mergers, which is again in agreement with the evolution of the tidal field for this case (cf. models PR and RR of Section~\\ref{sec:spin}). However, they noted a negative correlation between the pericenter distance and the starburst, in the sense that close passages correspond to \\emph{low} SFR, for mergers \\emph{only}. It is yet unclear to what extent this trend can be attributed to the response of the gas in the early stages of the merger. Because the statistics reported here do not account in details for the sequence of events that take place, we can not be specific about the likelihood of forming stars at a precise time. A clearer picture will emerge once high-resolution hydrodynamics is included in the dynamical scenarios that we have reported here. \\subsection{Age distribution function} In all merger simulations that we have carried out, a double exponential law describes well the compressive mode distribution in age, for both the longest uninterrupted sequence (\\emph{lus}) and the total time (\\emph{tt}). We note that the longest periods of the \\emph{lus} statistics correspond to orbits near the centre of the galaxies (in isolation as well as during the interaction) while the transient features that give rise to the shorter periods of the \\emph{lus} distribution were found in the merger-induced tidal structures (such as bridges and tails). In Section~\\ref{sec:history}, we showed that the history of the gravitational tidal field along individual orbits is very complex. Indeed, to compute the net effect of the tides on a star cluster over time via a statistical approach, one should combine \\emph{both} the \\emph{lus} and \\emph{tt} distributions. This is so because the tidal field changes character (compressive or extensive) repeatedly along most orbits (see Fig.~\\ref{fig:orbits_particles}). To illustrate this further, consider once more the case of the Antennae model. Roughly speaking, the statistics of compressive modes for that case can be split in two phases: a first phase in the time interval $\\Delta T_1 = ]0,50] \\Myr$, during which the compressive modes are mostly uninterrupted; and a second phase $\\Delta T_2 = [50,100] \\Myr$ during which the compressive modes are a collection of discrete episodes. Putting this together, a proxy for the distribution of time spent in compressive tidal modes might consist in taking the \\emph{lus} distribution over the time interval $\\Delta T_1$, followed by the \\emph{tt} statistics over the interval $\\Delta T_2$. Fig.~\\ref{fig:age} plots the \\emph{lus} and \\emph{tt} distributions (red and blue lines) showing the usual double exponential, together with the linear combination of the two described above (black curve). \\begin{figure} \\plotone{f15.eps} \\caption{The \\emph{lus} and \\emph{tt} distributions (red and blue lines, respectively) of compressive modes for the Antennae model. The distributions consist in double exponential laws fits to the short and long timescales ($\\Delta T_1$ and $\\Delta T_2$). The black line represents a combination of both cases, using the short timescale from the \\emph{lus} and the longer one from \\emph{tt} (see text for details). The green points with horizontal binning are HST data lifted from \\citet{Fall2005}.} \\label{fig:age} \\end{figure} If we take the bold step of identifying the duration of the modes (x-axis on Fig.~\\ref{fig:age}), with the age of stellar associations, then the vertical y-axis may be interpreted as the number of associations per age interval, $dN/d\\tau$. The black curve on Fig.~\\ref{fig:age} can be fitted with a $\\tau^{-1.08}$ power law ($\\sigma = 2\\times 10^{-4}$) over the interval $t \\in [10, 100] \\Myr$). It is striking that the duration of the compressive modes is very similar to the $\\tau^{-1}$ distribution derived from HST data by \\citet{Fall2005} for the age of young star clusters. While this does not constitute a full cause-to-effect chain of reasoning, the unexpected agreement does offer a strong hint that time-dependent tidal fields bear on the demographics of young clusters in galaxy mergers. \\subsection{Energy of compressive modes} To characterize statistically the strength of the tides, we plot in Fig~\\ref{fig:distrib} the distribution of the maximum eigenvalue of the tidal tensor for all particles, at selected times. As mentioned before, a negative value stands for a fully compressive mode. It is clear that, at any stage of the merger, the distribution can not be fitted by simple analytical functions (such as a Gaussian or power laws). Its high degree of asymmetry with respect to the origin mainly stems from the distribution of eigenvalues in the galaxies in isolation (black curve on Fig.~\\ref{fig:distrib}), where a large fraction of the mass ($\\sim 97\\%$) experiences extensive but weak tidal modes, while the few percent mass elements nearest to the galactic center experience compressive modes of large, negative eigenvalues. The distribution of compressive and extensive modes tends to become more even-handed in the course of the merge, in terms of the occupation percentage at given $|\\lambda|$. This is illustrated at the first pericenter passage, $t = 0 \\Myr$, and at $t = 300 \\Myr$ when the merger is in full swing. One can deduce from this that the strong destructive effect of the extensive tidal field acting at some point in the system is balanced out, on average, by a compressive mode of comparable strength in some other region, like e.g. the tidal tails. Given that individual orbits enter and exit compressive modes several times during the merger (cf. Fig.~\\ref{fig:orbits_particles}) one may anticipate that transitions from compressive to extensive modes are relatively sharp over a timescale of $\\sim 2 \\Myr$. \\begin{figure} \\plotone{f16.eps} \\caption{Normalized distribution (in percent) of the maximum eigenvalue ($\\lambda_\\mathrm{max}$) of the tidal tensor for the Antennae run. Three key steps are shown: the configuration in isolation ($t = -100 \\Myr$), the first pericenter passage ($t = 0 \\Myr$) and the merger stage ($t = 300 \\Myr$).} \\label{fig:distrib} \\end{figure} We already noted how the full impact of the compressive tides could only be quantified once the hydrodynamics of star formation is included in the modeling. Despite this caveat, one can estimate the relative importance of compressive tides by comparing the extra binding energy they add to a cluster, to the feedback from stellar winds. For simplicity, let us assume an isotropic tidal field, i.e. described by a tensor set equal to the identity matrix times $\\lambda$. For a mass distribution of total mass $M_\\mathrm{t}$, the corresponding binding energy driven by the tide is \\begin{equation} \\label{eqn:energy} E_\\mathrm{T, cluster} = -\\frac{1}{2} \\lambda \\int_0^{M_\\mathrm{t}}{r^2 \\ dm} = -\\frac{1}{2} \\lambda \\alpha \\ M_\\mathrm{t} R_\\mathrm{t}^2 \\end{equation} where $\\alpha$ is a dimensionless quantity of the order of unity, which encapsulates details of the mass distribution, and $R_\\mathrm{t}$ is the truncation radius. Applying Eq.~\\ref{eqn:energy} to the resolution of our simulations ($M_\\mathrm{t} = 10^5 \\msun$ and $R_\\mathrm{t} = 10^2 \\U{pc}$), we get a typical value for $\\lambda$ of $\\sim -10^{-30} \\U{s^{-2}}$ ($< 0 $ for a compressive tide). This yields an estimate for the tidal binding energy of $\\sim 10^{49} \\U{ergs}$. Clearly, this value does not compare to the energy release $\\sim 10^{51} \\U{ergs}$ of a single supernova explosion, and hence SN-driven gas expulsion would proceed largely unaffected by the tides. However, the binding energy computed from Eq.~\\ref{eqn:energy} would compare favorably to the kinetic energy of an O-star's wind \\citep{Cappa1998, Martin2008}. In that case the mass loss driven by the winds of early-type stars would be slowed down considerably. The enhanced mass retention would then help stop or reduce the dissolution of young clusters, an aspect that has immediate consequences to the question related to their mass functions in mergers. In addition to star clusters, compressive tidal modes might also be important for the formation of TDGs. Detailed high-resolution $N$-body simulations showed that the formation of TDGs is virtually impossible in purely stellar systems \\citep{Wetzstein2007}. Consequently, a dissipative component (gas) is a necessary ingredient to form TDGs. Because the compressive region will span kiloparsec-size volumes, enough stars may form that a TDG will ensue. The binding energy of the tidal field estimated from Eq.~\\ref{eqn:energy} for a mass $M_\\mathrm{t} = 10^9 \\msun$ and a size of $10 \\kpc$ (typical values for the compressive regions found in the outer parts of our merger simulations) is $\\sim 10^{57} \\U{ergs}$, or the kinetic energy release of $10^6$ type-II supernovae. Now, if one converts $10^9 \\msun$ of gas using a standard stellar initial mass function and a star formation efficiency (SFE) of $10-20 \\%$, one ends up with on order of a million type-II supernova candidate stars. While this calculation is rough, it does allow us to conclude that the energy reservoir of compressive tides must be factored in more detailed analysis of TDG formation models, as it is at least comparable to SN energy feedback. \\subsection{Perspectives} In this last, more speculative section we explore briefly three key areas for future work that will draw from the current study: \\begin{itemize} \\item The first and more obvious is the modeling of the time-evolution of cluster mass functions driven by galactic tides. A number of previous studies derived cluster dissolution rates based on steady external tides \\citep{Baumgardt2003, Gieles2008a, Gieles2008b, Kuepper2008, Vesperini2009, Ernst2009}. Other approaches encapsulate the tide's time-evolution in an approximate ``shock'' treatment as the model cluster orbits the host galaxy (\\citealt{Spitzer1987}; see \\citealt{Gieles2007} for a recent application to spiral galaxies). We note that the tidal tensor approach taken in this paper encompasses those mentioned here as limiting cases. But the important point to stress in our view is that the statistics of tidal fields change very significantly once we consider out-of-equilibrium dynamics. It will be important in future to take into account this evolution of the tidal field itself when considering the evolution of cluster mass distributions, especially in mergers. \\item Secondly, the tidal field bears down on star-forming regions at the mass- and length scales of star clusters in merging galaxies, and hence on the hydrodynamics of fragmentation and star formation in such systems. Future modeling of the first $\\sim 10^6 \\U{yr}$ of a giant molecular cloud embedded in a realistic, fully evolving background tidal field should shed new light on questions pertaining to the survival of young (open or globular) clusters \\citep{Bastian2006, Goodwin2008}. The survival of clusters is set from a combination of several effects, e.g. the expulsion of the residual gas \\citep{Goodwin1997, Geyer2001, Chen2008}; and two-body relaxation \\citep{Portegies2008}. How the rapid evolution of the tides that we have reported here changes the picture derived from the secular evolution of clusters (the weak field limit, see e.g. \\citealt{Goodwin1997, Boily2003a, Boily2003b, Baumgardt2007, Parmentier2009}) is an aspect worthy of further consideration. \\item Finally, several authors have confirmed in recent years the presence of multiple stellar main sequences (MS) in resolved massive clusters \\citep[see e.g][]{Piotto2007, Milone2009}. \\citet{dAntona2004} suggested that a stellar initial mass function (IMF) with a flat slope at the high-mass end would enhance CNO abundances yet leave relatively unchanged the abundances of iron-peak elements, providing a fit for the multiple threads in cluster HR diagrams. It is unclear why a bias in the stellar IMF should develop in some clusters. One possibility comes in the form of an age-spread as broad as $\\sim 10 \\Myr$, which could be interpreted as a second episode of star formation. This would open up the possibility that (i) clusters are self-enriched (by retaining slow stellar winds); or (ii) accrete gas from gas-rich environments met on their galactic orbit. In that context it is clear that compressive gravitational tides would help increase the likelihood of a second episode of star formation. Gas accretion by the cluster should impact on \\emph{both} the CNO abundances and those of the iron-peak elements, and so offers little hope of a viable explanation to this problem. Furthermore, the cluster has little time over $10 \\Myr$ to move significantly away from the volume where it first condensed before the second burst of star formation takes place. Thus, this scenario would preserve the homogeneity of the metal abundances the cluster already holds. Self-enrichment, on the other hand, should be more effective in broadening CNO abundances since the new abundances would remain correlated with those of the host cluster at birth. Recent modeling of rotating massive stars has shown that slow disk-like ejecta may be more effectively retained by the cluster than was assumed up to now \\citep{dErcole2008, Decressin2007}. Clearly the retention of light elements would be enhanced if the cluster was undergoing a compressive tidal mode at the time when such winds were operative. It is too early to pin down quantitatively the actual impact of tidal modes on such cluster evolutionary processes, which will surely require exhaustive modeling of the hydrodynamics and the chemical evolution, not done here. \\end{itemize}" }, "0910/0910.0475_arXiv.txt": { "abstract": "{% We summarize our optical monitoring program of VY Scl stars with the SMARTS telescopes, and triggered X-ray as well as optical observations after/during state transitions of V504 Cen and VY Scl. } ", "introduction": "VY Scl stars are a subclass of cataclysmic variables (CVs) that are optically bright most of the time, but occasionally drop in brightness by several magnitudes at irregular intervals (Bond 1980, Warner 1995, Honeycutt \\& Kafka 2004). The transitions between the brightness levels take place \\linebreak within days to weeks. In their high states, these variables have the largest time-averaged mass transfer rate $\\dot{M}$ (of the order of 10$^{-8}$ M$_{\\odot}$/yr) among CVs, and thus are thought to be steady accretors with hot disks. The cause of the transitions is widely debated. The conventional view is that a strong reduction (or even cessation) of the mass transfer rate, either due to the magnetic spot covering temporarily the $L_1$ region (Livio \\& Pringle 1994) or due to non-equili\\-brium effects in the irradiated atmosphere of the donor (Wu et al. 1995) causes the deep low states. A major problem of this scenario is the lack of dwarf nova (DN) outbursts during the low states (Leach et al. 1999): the disk remains subject to the thermal/viscous instability as it must drain its remaining gas onto the white dwarf (WD) through a series of DN eruptions. Another serious difficulty is that the observed dual-slope rises (Honeycutt \\& Kafka 2004), which are faster when the system is fainter, are opposite to the expected behaviour of an accretion disk since rebuilding the disk from an extended low state should initially take place on the slow viscous timescale of the low-state disk, eventually switching to a faster rise as the disk goes into DN outburst on the faster thermal timescale. Two unconventional views invoke quasi-stable burning of the accreted hydrogen on the white dwarf surface or a magnetic nature of the white dwarf. In the former case, VY Scl stars may be transient supersoft X-ray sources during optical low-states (Greiner \\& DiStefano 1999, Grei\\-ner \\linebreak 2000, Honeycutt 2001). Discovering more nearby supersoft X-ray sources is potentially important, since some of them may be progenitors of Type Ia supernovae; in addition, the sources may play an important role as ionizers of the interstellar medium. This picture is motivated by the behaviour of the classical supersoft X-ray source RX J0513.9--6951 which has quasi-periodic optical intensity dips of $\\sim$4 weeks duration (Southwell et al. 1996, Reinsch et al. 2000) simultaneously to X-ray high states. This is interpreted as being due to a drop in accretion rate which leads to a contraction of the WD, and in turn a hotter WD surface temperature. Besides this phenomenological similarity, there exist independent observational hints for transient supersoft X-ray emission in 5 VY Scl stars: V751 Cyg (Greiner et al. 1999), V~Sge (Greiner \\& Teeseling 1998), DW UMa (Knigge et al.\\ 2000), SW Sex (Groot et al.\\ 2001), BZ Cam (Greiner et al. 2001). This conjecture that VY Scl stars are the low-mass extension of supersoft X-ray binaries (SSB) can only be tested by combined optical and X-ray observations. If it can be pro\\-ven, it is expected to have far-reaching consequences. For example, if a relation to SSBs can be established, it would add a whole group of optically bright objects that are much easier to study than the optically-faint SSBs. Second, because the presence of WDs in SSBs has never been proven, a connection of SSBs and VY Scl stars would add support to the standard model of supersoft X-ray sources (van den Heuvel et al.\\ 1992). Third, the irradiation of the donors in supersoft X-ray sources is much stronger than in CVs (and VY Scl stars), and therefore the mechanism proposed by Wu et al.\\ (1995) for VY Scl stars could be readily applicable to SSBs. In the case of a magnetic primary, a low magnetic field of order 5$\\times$10$^{30}$ G cm$^3$ would disrupt the inner, otherwise unstable accretion disk, and thus prevent outbursts in low states, suggesting that VY Scl stars may all be intermediate polars (Hameury \\& Lasota 2002). While variable circular polarization (and consequently a magnetic field) has been found in two VY Scl stars (LS Peg and V795 Her), the majority has no measurable magnetic field. This conjecture can be \"easily\" tested by X-ray observations (searching for X-ray pulsations) or optical spectropolarimetry. Here, 10 yrs after the first suggestion (Greiner \\& DiStefano 1999), we report on our results of long-term optical monitoring of southern VY Scl stars, and triggered follow-up optical and X-ray observations when a VY Scl star went into an optical low-state. ", "conclusions": "For V504 Cen, we tested for a hot source in March 2006. The optical spectra do not show any significant sign of HeII emission, and Chandra and XMM ToO observations do not show a strong supersoft X-ray component. The magnetic option was tested with spectro-polarimetry, but no sign of a substantial magnetic field was found. It thus appears that the suspected H surface burning during the low-states in VY Scl stars either happens at very low temperatures (below 20 eV), or is not a generic feature in this class of objects. Similarly, a magnetic nature of the WD could not be proven. This indicates that in V504 Cen both scenarios proposed as ``unconventional'' views could not be substantiated, but also the conventional Lagrange point blockage model is challenged by the very long duration of the optical low-state." }, "0910/0910.5299_arXiv.txt": { "abstract": "{ The discovery of short-period Neptune-mass objects, now including the remarkable system HD69830 \\cite{lovis06} with three Neptune analogues, raises difficult questions about current formation models which may require a global treatment of the protoplanetary disc. Several formation scenarios have been proposed, where most combine the canonical oligarchic picture of core accretion with type I migration (e.g.~\\citealt*{terq}) and planetary atmosphere physics (e.g.~\\citealt{alib}). To date, due in part to the computational challenges involved, published studies have considered only a very small number of progenitors at late times. This leaves unaddressed important questions about the global viability of the models. We seek to determine whether the most natural model -- namely, taking the canonical oligarchic picture of core accretion and introducing type I migration -- can succeed in forming objects of 10 Earth masses and more in the innermost parts of the disc. } { This problem is investigated using both traditional semianalytic methods for modelling oligarchic growth as well as a new parallel multi-zone N-body code designed specifically for treating planetary formation problems with large dynamic range \\citep{mcn09}.} { We find that it is extremely difficult for oligarchic tidal migration models to reproduce the observed distribution. Even under many variations of the typical parameters, including cases in which after the amount of mass in our disc is greatly increased above the standard Hayashi minimum-mass model, we form no objects of mass greater than 8 Earth masses. By comparison, it is relatively straightforward to form icy super-Earths.} {We conclude that either the initial conditions of the protoplanetary discs in short-period Neptune systems were substantially different from the standard disc models we used, or there is important physics yet to be understood and included in models of the type we have presented here.} ", "introduction": "\\label{section:intro} At present, there are $\\sim\\!$19 known extrasolar planets\\footnote{Data taken from the Extrasolar Planets Encyclopedia, Schneider, J., http://exoplanet.eu} with estimated masses between 0.03 and 0.12 \\MJup, or between $\\sim\\!10$ and $\\sim\\!40$ \\ME. With one exception -- OGLE-05-169L b, at 2.8 AU -- all of the planets have semimajor axes smaller than 1 AU, and so are reasonably called hot Neptunes. In fact, with only one more exception, HD69830 d, all have semimajor axes $< 0.23$ AU. To determine whether the objects are genuine Neptunes, i.e.~ice giants, and not merely very large rocky bodies better thought of as super-Earths, knowledge of the mean densities is required. Three of these objects have known radii, and two have mean densities compatible with being a Neptune-like body. GJ436 b \\citep{butler2004} has a density of $\\sim\\!1.69\\,\\mathrm{g/cm^3}$ \\citep{torres}, and HAT-P-11 b \\citep{bakos} has a density of $\\sim\\!1.33\\,\\mathrm{g/cm^3}$; compare Neptune with $1.64\\,\\mathrm{g/cm^3}$. The HD69830 system \\citep{lovis06} is particularly interesting. It contains three Neptune-like objects: HD69830 b at 0.0785 AU, of 10.5 \\ME$\\,$ and eccentricity 0.1; HD69380 c at 0.186 AU, of 12.1 \\ME$\\,$ and eccentricity 0.13; and HD69380 d at 0.63 AU, of 18.4 \\ME, and eccentricity 0.07 ($\\pm0.07$, so a near-circular orbit is possible). The system was also reported to have a disc of warm infrared-emitting dust located between planets c and d \\citep{beichman2005}, presumably due to a collisionally active remnant asteroid belt. More recent work \\citep{lisse07} suggests instead that the infrared excess is due to a debris disc outside the known planets at $\\simeq 1$ AU, probably resulting from the breakup of an asteroid. This system clearly provides a rigorous test of planet formation and migration theories. Indeed, short-period Neptune-mass bodies have several advantages as probes of planetary origins. They are large enough to be observable, but small enough that we need not consider gravitational disc instability as a formation process. They are also large enough to undergo significant type I migration, but not so large that they can open a gap in the disc and undergo type II migration \\citep{pap84, bryden99, crida06}. Accordingly, hot Neptunes can yield less ambiguous tests of type I migration than more massive planets which could have significantly perturbed the gas disc they were embedded in, and their existence provides strong evidence that some kind of type I migration is operating in protoplanetary discs. Hot Neptunes are therefore useful for exploring the intersection of the classical oligarchic core accretion picture (\\citealt*{ko98}; e.g.~\\citealt*{cham01, thommes03}) with type I migration \\citep{ward97}. Two main classes of scenario exist in the literature: one which concentrates on the atmospheric gas physics, and one which concentrates on the dynamics of the protoplanetary interactions. \\cite{alib} incorporate sophisticated atmospheric physics and follow the evolution of effectively isolated cores through the disc as they grow via planetesimal accretion, migrate inwards due to type I effects, accrete gas after the accretion rate of solids drops, and finally have their atmospheric mass reduced after arriving in the short-period region by evaporation. However, the history for the HD69830 system proposed in \\cite{alib} is difficult to reconcile with the oligarchic paradigm \\citep{ko98}. Their best-fitting model involves exactly three seed objects of masses $M = 0.6\\,\\ME$ at semimajor axes 3 AU, 6.5 AU, and 8 AU, which migrate to small semimajor axis through a mostly pristine planetesimal disc and therefore can accrete a fair amount of material \\citep{growthofmig}. Where did these three progenitor objects come from? In an oligarchic framework, accretion in a narrow region -- even in the presence of type I drag (see \\citealt{komI, mcn05, komII}) -- results in many roughly equal-mass bodies separated by a distance of $10\\sim20$ Hill radii. These bodies then interact and merge in the giant-impact phase of planet formation, and consume the accessible remnant planetesimal disc. That is, if a half-Earth-mass seed can successfully form at 3 AU, there should be a large number of similar seeds of varying mass inside, each of which will accrete and migrate in similar ways. Moreover, the interior seeds are likely to have formed first. The net result is that any such inward-migrating seed should be migrating through a region highly depleted by the previous seeds, and by the time an 0.6 \\ME$\\,$ core is formed at 3 AU much of the interior disc should have gone to completion, and possibly formed its own planets \\citep{cham08}. There is no obvious mechanism to suppress embryo growth everywhere in the disc except at three specific locations. An additional problem is that simulations have demonstrated that cores migrating through a planetesimal swarm are unable to grow at the rate prescribed by the \\cite{alib} model. In the presence of gas drag, mean motion resonances cause the majority of the interior planetesimals to be shepherded rather than being accreted, resulting in planets whose masses are too low, and in the wrong ratio, compared to those observed in the HD69830 system \\citep{payne2009}. Though not aiming at HD69830 in particular, \\cite{terq} study the formation of hot super-Earths and Neptunes by following the evolution of $10-25$ planets of $0.1$ or $1\\,\\ME$ placed interior to 2 AU under type I drag, include tidal interactions with the star, and use an inner cavity in the gas disc at $\\sim\\!0.05$ AU. For various disc parameters, they succeed in making several objects of mass $\\geq 8 \\,\\ME$, with a maximum of 12 \\ME. They find that the typical result has between 2--5 planets, usually on near-commensurable orbits (with strict commensurability often being broken by tidal circularization). It is not clear whether this will scale up to larger masses, as they performed only one run with total mass larger than 12 \\ME (namely 25 \\ME), and so an HD69830-like system would not appear in their results even if the model would have succeeded in producing them. Very little material was lost from their runs, suggesting it might be possible. However, their initial configurations are difficult to reconcile with an oligarchic migration process. Scenarios involving oligarchic formation and type I migration do not exhaust the possible formation histories of the low-mass hot exoplanet population, although they are arguably the most natural. \\cite{raymond08} surveys several other proposed possibilities (for the terrestrial-mass regime): in situ formation; shepherding by migrating giant planets or secular resonance sweeping; tidal circularization; and photoevaporation of giant planets. We will concentrate instead on the simplest oligarchic type I migration picture, which should have more success self-consistently generating an icy planet population, and attempt to determine whether the fiducial models can reproduce the observed distribution of hot Neptunes. If they can, all the better; if they cannot, then the specifics of their failure may point in the direction of a solution and help us choose between the various possibilities on offer. To understand how short-period Neptunes are formed, we need to move toward self-consistent global N-body models such as those which have proved useful in understanding the formation of the terrestrial planets and the outer solar system. One problem presents itself immediately: the short-period exoplanet problem has a formation time to dynamical time ratio $\\sim\\!30$ times larger than the equivalent terrestrial formation problem, and $\\sim\\!1000$ times that of the equivalent Jovian core formation problem. The lifetime of the gas disc (with a probable upper limit of $\\sim6$ Myr) is a common reference time-scale for each problem, as gas giants must form while there is still gas for them to accrete, and even ice giants which attain short-period orbits must have migrated while there is gas present. However, the characteristic orbital periods for each formation problem vary from $0.01$ yr at 0.05 AU to $0.35$ yr at 0.5 AU to $11$ yr at 5.0 AU. (Indeed, $0.01$ yr is optimistic, as many exoplanets are closer to their parent stars than 0.05 AU.) Since most state-of-the-art planetary N-body codes require the integration timestep to be some small fraction of the orbital period of the innermost object, preserving the same wall-clock run times that researchers have become accustomed to would usually require reducing the number of planetesimals in a simulation by orders of magnitude, producing an unacceptable decrease in resolution. It is extremely unlikely that this difficulty can be eliminated entirely, as it is a consequence of the strong dependence of orbital period on semimajor axis in the Kepler problem. That said, the challenge can be managed to some extent by taking advantage of the scale separation caused by the troublesome dynamic range. The standard approaches use a common drift timestep for all particles (and therefore ``over-integrate'' the more distant objects) and compute the (non-encountering) forces between the particles at the same frequency, both of which involve far more computation on the distant, slow-moving objects than is necessary to preserve qualitatively accurate dynamics. In \\cite{mcn09} the authors combine various techniques in the literature (\\citealt*{dll98, cham99, saha92}) to construct a new algorithm which allows for radial zones with different timesteps and different inter-zone force evaluation frequencies, but reduces to the proven techniques of \\cite{dll98} and \\cite{cham99} for objects within the same zone. This allows new trade-offs between force accuracy and run time. In our first paper, \\cite{mcn09}, we addressed the numerical challenges of studying oligarchic models of short-period exoplanet formation. Here we apply a new code with a parallel implementation of those methods, to determine the ``reference population'' of planets predicted by the fiducial models. We integrate global planetesimal discs extending from 0.05 AU to 10 AU under various models of the protoplanetary disc. Our models, both semianalytic and N-body, are described in \\S\\ref{modeldescr}, and the simulation conditions in \\S\\ref{simcond}. Simulation results are presented in \\S\\ref{results} and discussed in \\S\\ref{discussion}, and we conclude in \\S\\ref{section:conc}. ", "conclusions": "\\label{section:conc} We performed 48 simulations of various oligarchic migration scenarios to determine whether the simplest standard approach can succeed in forming a population of short-period Neptune systems under common assumptions for the protoplanetary disc parameters. Multiple numerical techniques were applied: semianalytic techniques for the first 0.4 Myr, our new parallel multizone N-body code for the accretion phase while gas is present up to 6 Myr, and the more traditional SyMBA approach for the late stage to 100 Myr. We find that over a wide range of disc conditions, it is difficult to form planets of mass greater than $3-4 \\ME$. Our most successful runs involved $\\sim$ 5 times the mass of the MMSN, surface density varying as $r^{-1/2}$, a disc decay time-scale of 1 Myr, and a migration efficiency of 0.3. Our most common planet outcomes are of Earth-mass objects, with the terrestrial planets having ice fractions from 0.0 to 0.75 (the maximum possible in our simulations). The larger objects have higher ice fractions, with the median being 0.60 for objects above 1 \\ME, and 0.36 below. In none of the cases did we succeed in forming an object of greater than 7.5 \\ME inside 2 AU, much less inside 0.5 AU, and the total embryo mass remaining inside 2 AU was always less than 17 \\ME. The existence of an upper limit and the weak dependence on most parameters is in accordance with the predictions of \\cite{komII}, and in rough agreement with the predictions of \\cite{cham08} except at large disc masses. Nevertheless, we should be wary of making predictions based on these results regarding extrasolar planetary systems, as they entirely fail to reproduce the short-period Neptune planetary population that we know exists. Our failure can be compared to several previous successes in the literature, which either (1) adopt initial conditions which are not easily reconciled with an oligarchic growth picture, (2) use an inner edge to the migration (which is defensible but will have difficulty explaining more distant Neptunes), or (3) neglect inter-embryo dynamics and use an embryo merger condition which is calibrated to an effective inter-embryo separation (10 Hill radii) which is considerably smaller than we observe. Varying parameters which we kept constant such as the gas-to-dust ratio, incorporating additional accretion physics such as fragmentation, moving to extremely large disc masses or extremely weak migration, and simply performing more runs (and hoping for fortuitous late mergers), could possibly succeed in improving the maximum mass reached by a factor of two and therefore into the Neptune-like region. It seems quite unlikely that they will increase the median mass enough to comfortably produce a population of multiple-planet short-period Neptune systems. We conclude that forming a system like HD69830 will probably require a significant revision to the simple models explored here. If the standard oligarchy-plus-type-I-migration picture fails to reproduce the observed distribution of short-period exoplanets even at more extreme parameter values, then we must consider non-standard models. Oligarchy is relatively well understood both analytically and numerically; by comparison type I migration is sensitive to poorly understood properties of the gas disc such as disc turbulence and local thermodynamic time-scales. In a follow-up paper we consider the implications for hot exoplanet formation via oligarchy of alternate migration models (such as \\citealt*{paard06}) which show some promise." }, "0910/0910.0643_arXiv.txt": { "abstract": "% The {\\sl F-GAMMA}-project is the coordinated effort of several observatories to understand the AGN phenomenon and specifically blazars via multi-frequency monitoring in collaboration with the {\\sl Fermi}-GST satellite since January 2007. The core observatories are: the Effelsberg 100-m, the IRAM 30-m and the OVRO 40-m telescope covering the range between 2.6 and 270 GHz. Effelsberg and IRAM stations do a monthly monitoring of the cm to sub-mm radio spectra of 60 selected blazars whereas the OVRO telescope is observing roughly 1200 objects at 15 GHz with a dense sampling of 2 points per week. The calibration uncertainty even at high frequencies, is of a few percent. 47\\% of the Effelsberg/Pico Veleta sample is included in the LBAS list. An update of the monitored sample is currently underway. ", "introduction": "The {\\sl unified model} paradigm for the Active Galactic Nuclei (AGN), attributes the wealth of AGN classes mostly to different parameters of the same principal system viewed at different angles of the line-of-sight to the relativistic jet. Radiogalaxies and blazars are two extremes: in the former the jet is seen from large whereas in the latter from small angles ($\\theta\\le 20^\\circ$). Blazars are dominated by relativistic beaming and they exhibit intense variability at all wavelengths and time scales, highly superluminal apparent speeds, a significant degree of fractional polarization and polarization variability. Shock-in-jet \\citep[][]{marscher1985, hughes1985, marscher1996}, colliding relativistic plasma shells \\citep{guetta2004}, lighthouse effect \\citep{camenzind1992} or MHD-instabilities in the accretion disks \\citep{begelman1980,villata1999} are some of the models that attempt to explain variability at different time scales. Although the exact physical processes at play are unclear, the study of the temporal behavior of the SED can shed light on emission mechanism since different mechanisms predict different variability patterns \\citep[e.g.][]{bottcher2002}. Hence, multi-frequency monitoring of blazars is essential in understanding the blazar physics. ", "conclusions": "{\\bf The F-GAMMA alliance: }The Large Area Telescope (LAT) on-board {\\sl Fermi}-GST offers a unique opportunity covering the $4\\pi$ sky every three hours providing $\\gamma$-ray light curves at $\\sim$30 MeV - 300 GeV. To exploit that, we have initiated a tightly coordinated ``alliance'' of teams \\citep[F-GAMMA project,][]{angelakis2008, fuhrmann2007} in order to understand the AGN physics via multi-frequency monitoring. The core program involves the 100-m MPIfR telescope (Effelsberg, Bonn, 2.64 - 43.00 GHz), the 30-m IRAM telescope (Pico Veleta, Granada, 86, 142, 220 and 270 GHz) and the 40-m OVRO telescope (Owens Valley, California, 15 GHz). The Effelsberg and the IRAM telescopes do a monthly monitoring of $\\sim$60 blazars compiled on the basis of their $\\gamma$-ray detectability \\citep[EGRET][]{hartman1999} and their presence in the ``high priority list'' of the LAT AGN group. The OVRO telescope is monitoring a statistically complete sample of 1200 sources based on the CGRaBS catalog \\citep{healy2008}. The sampling rate is 2-3 times a week \\citep[see][]{richards2009,maxmoerbeck2009}. \\\\{\\bf The first 2.5 years:} The F-GAMMA project has been running since January 2007 at Effelsberg and June 2007 at OVRO and Pico Veleta. The data products of the first 2.5 years can be publicly accessed at {\\tt www.mpifr.de/div/vlbi/fgamma}. The cross-correlation of the Effelsberg/Pico sample with the LBAS list \\citep{abdo2009} showed that 47\\% of our sources are detected by {\\sl Fermi}-GST. A $\\chi^2$ test shows that practically all our sources are variable at all frequencies at a confidence level of 99.9\\,\\%. The variability amplitude (as parameterized by the modulation index $m=rms/$ ) increases with frequency as expected from frequency flare evolution arguments. For Effelsberg, the calibration uncertainties are kept to the level of $3-5\\,\\%$ at high frequencies ($\\nu\\ge 23.05$\\, GHz) and of $0.5 - 1\\,\\%$ at the low frequencies ($\\nu<23.05$\\, GHz). The characteristic time scales present in the light curves vary from weeks to years while there is a wealth of variability patterns both in the lightcurves and in the spectra. The findings are summarized in Fuhrmann et al. (in prep.), Richards et al. (in prep.) and Angelakis et al. (in prep.). Given the limited fraction of our sources in the LBAS list we are currently compiling an updated sample." }, "0910/0910.5120_arXiv.txt": { "abstract": "We describe the spectroscopic target selection for the Galaxy And Mass Assembly (GAMA) survey. The input catalogue is drawn from the Sloan Digital Sky Survey (SDSS) and UKIRT Infrared Deep Sky Survey (UKIDSS). The initial aim is to measure redshifts for galaxies in three $4\\times12$ degree regions at 9\\,h, 12\\,h and 14.5\\,h, on the celestial equator, with magnitude selections $r<19.4$, $z<18.2$ and $K_{\\rm AB}<17.6$ over all three regions, and $r<19.8$ in the 12-h region. The target density is $1080\\,\\persqdeg$ in the 12-h region and $720\\,\\persqdeg$ in the other regions. The average GAMA target density and area are compared with completed and ongoing galaxy redshift surveys. The GAMA survey implements a highly complete star-galaxy separation that jointly uses an intensity-profile separator ($\\dsg=r_{\\rm psf}-r_{\\rm model}$ as per the SDSS) and a colour separator. The colour separator is defined as $\\dsgjk=J-K-f(g-i)$, where $f(g-i)$ is a quadratic fit to the $J-K$ colour of the stellar locus over the range $0.30.2$. From two years out of a three-year AAOmega program on the Anglo-Australian Telescope, we have obtained 79\\,599 unique galaxy redshifts. Previously known redshifts in the GAMA region bring the total up to 98\\,497. The median galaxy redshift is 0.2 with 99\\% at $z<0.5$. We present some of the global statistical properties of the survey, including \\newtext{$K$-band galaxy counts}, colour-redshift relations and preliminary $n(z)$. ", "introduction": "\\label{sec:intro} Galaxy redshift surveys provide a fundamental resource for studies of galaxy evolution. The redshift of a galaxy can be used to obtain a distance assuming a set of cosmological parameters, modulo peculiar velocities, and a well-defined selection function enables the comoving number density of galaxies to be estimated as a function of various properties, e.g., galaxy luminosity functions \\citep{Schechter76,BST88,norberg02,blanton03ld}. In addition, using the combined sky distribution and distance information, the clustering properties of galaxies can be determined \\citep{DGH78,dLGH88,norberg02clus,zehavi05} and the velocity dispersion of galaxies in groups and clusters can be used to infer dark-matter halo masses \\citep{Zwicky37,HG82,MFW93,carlberg96,eke04,berlind06}. The target selection algorithm and area covered by a redshift survey relate to the redshift range and volume surveyed. The industry of these surveys started in the 1980's with surveys of $\\sim2500$ galaxies over large sky areas \\citep{davis82,saunders90} and a deeper survey of 330 galaxies over $70\\,\\sqdeg$ \\citep{peterson86}. It expanded and diversified in the 1990's with surveys such as the wide-but-shallow CfA2 redshift survey, Las Campanas Redshift Survey, ESO Slice Project, and the deep-but-narrow Canada-France Redshift Survey. Figure~\\ref{fig:compare-grs} shows the surface density of galaxy spectra versus area for these and other surveys, and Table~\\ref{tab:z-survey-list} gives selections and references. The target density is a wavelength-independent metric for depth, at least for high-completeness magnitude-limited surveys. The advent of multi-object spectrographs such as the Two-Degree Field (2dF; \\citealt{lewis02df}) and Sloan Digital Sky Survey (SDSS; \\citealt{york00}) telescope have enabled redshift surveys of $>10^5$ galaxies: the 2dF Galaxy Redshift Survey (2dFGRS) and SDSS Main Galaxy Sample (MGS). \\begin{figure} \\centerline{ \\includegraphics[width=\\singlecolsize\\textwidth]{f01.ps} } \\caption{Comparison between field galaxy surveys with spectroscopic redshifts: {\\it squares} represent predominantly magnitude-limited surveys; {\\it circles} represent surveys involving colour cuts for photometric redshift selection; while {\\it triangles} represent highly targeted surveys. The colours represent different principal wavelength selections as in the legend. Filled symbols represent completed surveys. See Table~\\ref{tab:z-survey-list} for survey names and references.} \\label{fig:compare-grs} \\end{figure} \\begin{table*} \\caption{List of field galaxy redshift surveys. The surveys shown in Fig.~\\ref{fig:compare-grs} are listed in order of increasing area. They are mostly magnitude limited galaxy samples except for some with colour selection (CS). The information was obtained from the references and the survey websites.} \\label{tab:z-survey-list} \\begin{tabular}{lllcl} \\hline abbrev. & survey name & selection(s) & area/$\\sqdeg$ & reference \\\\ \\hline CFRS & Canada-France Redshift Survey & $I_{\\rm AB}<22.5$ & $0.14$ & \\citealt{lilly95} \\\\ LBG-z3 & Lyman Break Galaxies at $z\\sim3$ Survey & $R_{AB}<25.5$ with CS$^a$ & $0.38$ & \\citealt{steidel03} \\\\ VVDS-deep& VIMOS VLT Deep Survey deep sample & $I_{\\rm AB}<24.0$ & $0.5$ & \\citealt{lefevre05} \\\\ CNOC2 & Canadian Network for Obs.\\ Cosmology 2 \\rlap{...} & $R<21.5$ & $1.5$ & \\citealt{yee00} \\\\ zCOSMOS & Redshifts for the Cosmic Evolution Survey & $I_{\\rm AB}<22.5$, $I_{\\rm AB}\\la24$ with CS$^b$ & $1.7$ & \\citealt{lilly07} \\\\ DEEP2 & Deep Evolutionary Exploratory Probe 2 \\rlap{...} & $R_{\\rm AB}<24.1$ with CS$^c$ & $2.8$ & \\citealt{davis03} \\\\ Autofib & Autofib Redshift Survey & $b_{J}<22.0$ & $5.5$ & \\citealt{ellis96} \\\\ H-AAO & Hawaii+AAO K-band Redshift Survey & $K<15.0$ & $8.2$ & \\citealt{huang03} \\\\ AGES & AGN and Galaxy Evolution Survey & incl.\\ $R<20.0$, $B_{\\rm W}<20.5$ & $9.3$ & \\citealt{watson09} \\\\ VVDS-wide& VIMOS VLT Deep Survey wide sample & $I_{\\rm AB}<22.5$ & $12.0$ & \\citealt{garilli08}\\\\ ESP & ESO Slice Project & $b_{J}<19.4$ & $23.3$ & \\citealt{vettolani97}\\\\ MGC & Millennium Galaxy Catalogue & $B<20.0$ & $37.5$ & \\citealt{liske03}\\\\ {\\bf GAMA}&Galaxy And Mass Assembly Survey & $r<19.8$, $z<18.2$, $K_{\\rm AB}<17.6$ & $144$ & --- this paper --- \\\\ 2SLAQ-lrg& 2SLAQ Luminous Red Galaxy Survey & $i<19.8$ with CS$^d$ & $180$ & \\citealt{cannon06}\\\\ SDSS-s82 & SDSS Stripe 82 surveys & incl.\\ $u\\la20$, $r<19.5$ with CS$^e$ & $275$ & \\citealt{sdssDR4}\\\\ LCRS & Las Campanas Redshift Survey & $R<17.5$ & $700$ & \\citealt{shectman96}\\\\ WiggleZ & WiggleZ Dark Energy Survey & ${\\rm NUV}<22.8$ with CS$^f$ & $1000$ & \\citealt{drinkwater09}\\\\ 2dFGRS & 2dF Galaxy Redshift Survey & $b_{J}<19.4$ & $1500$ & \\citealt{colless01}\\\\ DURS & Durham-UKST Redshift Survey & $b_{J}<17.0$ & $1500$ & \\citealt{ratcliffe96}\\\\ SAPM & Stromlo-APM Redshift Survey & $b_{J}<17.1$ (1 in 20 sampling) & $4300$ & \\citealt{loveday92}\\\\ SSRS2 & Southern Sky Redshift Survey 2 & $B < 15.5$ & $5500$ & \\citealt{dacosta98}\\\\ SDSS-mgs & SDSS Main Galaxy Sample & $r<17.8$ & $8000$ & \\citealt{strauss02}\\\\ SDSS-lrg & SDSS Luminous Red Galaxy Survey & $r<19.5$ with CS$^g$ & $8000$ & \\citealt{eisenstein01} \\\\ 6dFGS & 6dF Galaxy Survey & $K<12.7$, $b_{J},r_F,J,H$ limits & $17000$ & \\citealt{jones09} \\\\ CfA2 & Center for Astrophysics 2 Redshift Survey & $B<15.5$ & $17000$ & \\citealt{falco99}$^h$ \\\\ PSCz & IRAS Point Source Catalog Redshift Survey & $60\\mu m_{\\rm AB}<9.5$ & 34000 & \\citealt{saunders00} \\\\ 2MRS & 2MASS Redshift Survey & $K<12.2$ & 37000 & \\citealt{erdogdu06} \\\\ \\hline \\end{tabular} \\begin{flushleft} Notes: $^a$CS by $U$-band `dropouts' for photometric redshifts ($z_{\\rm ph}$) $\\sim2.5$--3.5; $^b$CS for $z_{\\rm ph}\\sim1.4$--3.0, deeper limit over $1\\,\\sqdeg$; $^c$CS for $z_{\\rm ph}\\ga0.7$; $^d$CS for $z_{\\rm ph}\\sim0.45$--0.8; $^e$CS for $z_{\\rm ph}\\la0.15$; $^f$CS by ${\\rm FUV-NUV}>1.5$ (GALEX bands) and $20.50.2$) and strongly resolved (selected regardless of $\\dsgjk$) in SDSS $r$-band imaging. The thin grey line histograms represent unresolved samples in 36 randomly selected $1\\degr \\times 1\\degr$ regions.}} \\label{fig:test-auto-jk} \\end{figure} \\subsection{Magnitude limits} \\label{sec:mag-limits} The main scientific goal of GAMA that drives the choice of the minimum width of the survey geometry, and the magnitude selection, is the measurement of the halo mass function \\citep{driver09}. We chose $r$-band selection because it is most directly correlated with spectral S/N obtained (the filter falls in the middle range of the spectrograph). This ensures a high redshift success rate for a given target density. The $r$-band limits were chosen to give an average target density up to an order of magnitude higher than the SDSS MGS ($90\\,\\persqdeg$) and 2dFGRS ($140\\,\\persqdeg$). Given the limitations of efficient observing over two or three lunations each year, three fields were chosen covering 6 hours in RA. We compromised between area and depth by choosing a limit of $r<19.4$ in G09 and G15 ($670\\,\\persqdeg$), and $r<19.8$ in G12 ($1070\\,\\persqdeg$). These were defined using Petrosian magnitudes, following the strategy of the SDSS MGS. In consideration of measuring the stellar mass function, we included a near-IR selection using SDSS $z$-band and UKIDSS $K$-band. To ensure reliability and reasonable redshift success rate, these were also constrained by an $r$-band selection ($r_{\\rm model}<20.5$). The choice of SDSS model magnitudes rather than Petrosian is a consequence of the noise statistics. For Petrosian magnitudes, the noise is well behaved to $r\\simeq20$ \\citep{stoughton02}, while for fainter objects the model magnitudes are more reliable. Figure~\\ref{fig:mag-test} shows the pipeline-output magnitude errors versus magnitude. Also, the $K$-band selection was based on \\textsc{auto} magnitudes, and both \\textsc{auto} and model magnitudes use elliptical apertures. The additional selections were a small sample to $z_{\\rm model}<18.2$ and a sample to $K_{\\rm AB,auto}<17.6$. \\begin{figure} \\centerline{ \\includegraphics[width=\\smallcolsize\\textwidth]{f09.ps} } \\caption{Magnitude errors versus magnitude. The solid lines show the median errors obtained from the SDSS catalogue, with the regions representing the inter-quartile range (top for Petrosian, lower for model magnitudes). The vertical dash-dotted lines represent the $r$-band limits used in this paper (19.4 and 19.8 using Petrosian, and 20.5 using model magnitudes).} \\label{fig:mag-test} \\end{figure} Within the GAMA regions, the main survey selections are given by: \\begin{equation} \\begin{array}{ll} r_{\\rm petro}<19.4 & \\mbox{\\textsc{or}} \\\\ r_{\\rm petro}<19.8 \\mbox{~~\\textsc{and}~ in the G12 area} & \\mbox{\\textsc{or}} \\\\ z_{\\rm model}< 18.2 \\mbox{~~\\textsc{and}~~} r_{\\rm model} < 20.5 & \\mbox{\\textsc{or}} \\\\ K_{\\rm AB,auto} < 17.6 \\mbox{~~\\textsc{and}~~} r_{\\rm model} < 20.5 \\: . \\end{array} \\label{eqn:mag-limits} \\end{equation} Including the near-IR selections increases the G12 target density marginally (to $1080\\,\\persqdeg$) while increasing the G09 and G15 target density to $720\\,\\persqdeg$. Figure~\\ref{fig:color-bias} shows the colour bias for the near-IR selections. The $z$-band selection is complete to $(r-z)_{\\rm model} < 2.3$ at the faint limit, while the $K$-band selection is complete to $r_{\\rm model} - K_{\\rm AB,auto} < 2.9$ at the faint limit. A $z_{\\rm model}<18.2$ selection is formally missing 0.3\\% of objects because of the $r_{\\rm model}$ limit, while a $K_{\\rm AB,auto}<17.6$ selection is formally missing about 1\\% of objects. This is after applying star-galaxy separation. However, only very red objects are missed, which are more likely to be stars in spite of the star-galaxy separation or have incorrectly measured colours caused by mismatched apertures in the case of $r-K$ (the practical impact of these joint limits is discussed later in \\S~\\ref{sec:z-distribution}). \\newtext{See also Fig.~\\ref{fig:K-counts}, the $r<20.5$ limit only makes an obvious impact in the galaxy number counts at $K_{\\rm AB,auto}>17.8$ that is above our selection limit.} \\begin{figure} \\centerline{ \\includegraphics[width=\\singlecolsize\\textwidth]{f10.ps} } \\caption{Colour versus magnitude distribution for a near-IR sample. The black contours and points represent potential galaxy targets. The blue dashed lines show the limits imposed by our selection including the constraint $r_{\\rm model}<20.5$. The green lines show $r=19.4$ and 19.8 limits. The red contours represent most of the additional targets not selected by the $r_{\\rm petro}$ limits. These contours extend below the solid green line because of differences between Petrosian and model magnitudes.} \\label{fig:color-bias} \\end{figure} \\subsection{Masking} \\label{sec:masking} In order to avoid targeting galaxies with bad photometry because they are near bright stars or satellite trails, an explicit mask was constructed. The bright-stars mask was based on stars down to $V<12$ in the Tycho~2, Tycho~1 and Hipparcos catalogues. For each star, a scattered-light radius ($\\rscat$) was estimated based on the circular region over which the star flux per pixel is greater than 5 times the sky noise level. For each potential target, a mask parameter was defined as follows \\begin{equation} \\begin{array}{l} \\mbox{{\\small MASK\\_IC\\_12}} = 1 \\\\ \\mbox{{\\small MASK\\_IC\\_12}} = \\rscat/d \\\\ \\mbox{{\\small MASK\\_IC\\_12}} = 0 \\end{array} \\mbox{~for~} \\begin{array}{l} d \\le \\rscat \\\\ \\rscat < d \\le 5 \\rscat \\\\ d > 5 \\rscat \\end{array} \\end{equation} where $d$ is the distance to a $V<12$ star with radius $\\rscat$. In other words, the {\\small MASK\\_IC\\_12} value decreases from unity when $d \\le \\rscat$ to 0.2 when $d = 5 \\rscat$. A similar mask parameter {\\small MASK\\_IC\\_10} was defined using only $V<10$ stars. In addition, objects within an SDSS-database mask for holes, satellite trails and bleeding pixels had these mask values set to unity. After testing, we chose to select only objects with {\\small MASK\\_IC\\_10} $< 0.5$ and {\\small MASK\\_IC\\_12} $< 0.8$. The largest masked areas are shown in Fig.~\\ref{fig:ukidss-coverage}. These are between 0.01 and $0.07\\,\\sqdeg$ each. Most of the separate masked areas are significantly smaller ($<0.001\\,\\sqdeg$ or $<1'$ in radius). Overall, the total masked area is about $1.0\\,\\sqdeg$ and the unmasked area of the survey is estimated to be $143.0\\,\\sqdeg$. The mask was insufficient to remove all or nearly all objects with bad photometry. Therefore, as per SDSS selection, objects were selected to be \\textsc{not satur} from the \\textsc{flags} column in the \\textsc{PhotoObj} table. This basically excludes deblends of bright stars but will also reject galaxies that are blended with saturated stars. These however are likely to have bad photometry and falsely bright magnitudes. The stars causing this saturation, not accounted for by the Tycho mask, are probably around $V\\sim13$. The saturated-flag masking is not ideal. This is particularly the case for large nearby galaxies for which the angular size of the galaxy is a significant factor in determining the excluded sky area. In other words, the probability of a large galaxy having \\textsc{satur} set depends primarily on its size rather than the area of the diffracted and scattered light around stars. To increase the completeness of the input catalogue for large galaxies, exceptions for the mask and not-saturated criteria were made for galaxies from the Uppsala General Catalog (UGC; \\citealt{cotton99}) and Updated Zwicky Catalog (UZC; \\citealt{falco99}). In addition, exceptions to the not-saturated criteria were made for a selection of visually inspected galaxies that have \\textsc{not satur\\_center}. There are only 86 objects with an exception flag set (selected as part of the visual classification process described in \\S~\\ref{sec:vis-class}). In summary, the criteria for including objects is given by: \\begin{equation} \\begin{array}{l} (\\mbox{\\textsc{mask\\_ic\\_10}} < 0.5 \\mbox{~~\\textsc{and}~~} \\mbox{\\textsc{mask\\_ic\\_12}} < 0.8 \\mbox{~~\\textsc{and}} \\\\ \\mbox{\\textsc{not satur}} ) \\mbox{~~\\textsc{or}~~} \\mbox{the exception flag is set.} \\end{array} \\label{eqn:masking} \\end{equation} \\subsection{Surface brightness limits} \\label{sec:sb-limits} In addition to the implicit surface brightness (SB) limits from star-galaxy separation and detection, an explicit SB limit was applied given by \\begin{equation} 15.0 < \\effsb < 26.0 \\label{eqn:sb-limits} \\end{equation} where $\\effsb$ is the effective SB in ${\\rm mag\\,arcsec}^{-2}$ within the 50\\% light radius in the $r$-band (eq.~5 of \\citealt{strauss02}). Anything of lower SB is very likely to be an artifact, and anything of higher SB is a star. Figure~\\ref{fig:sb-sep} shows the distribution of objects in $r_{\\rm fibre}$ versus $\\effsb$ for GAMA main-survey targets. The lower limit of 15.0 does remove some objects, probably stars, not rejected by the masking or star-galaxy separation criteria (Eq.~\\ref{eqn:star-gal-summary}). The limit for $\\effsb$ of $26.0$ is 1.5~magnitudes deeper than the SDSS MGS cut, and is the point at which most of the objects are clearly artifacts. Note that the SDSS photometric pipeline is not complete for $\\effsb > 23$ (figs.~2--3 of \\citealt{blanton05}). Additional low-SB candidates could be recovered by searching coadded $g$, $r$ and $i$ images \\citep{kniazev04}. Nevertheless without deeper imaging, the data will remain incomplete at low SB well before our explicit limit. \\begin{figure*} \\includegraphics[width=\\middlecolsize\\textwidth]{f11.ps} \\caption{Bivariate distribution of $r_{\\rm fibre}$ versus $\\effsb$. The black contours and points represent objects that are not masked and pass star-galaxy separation. The grey lines outline the selection limits: $15.0<\\effsb<26.0$ is the restriction for the science catalogue; while objects with fibre magnitudes fainter than 22.5 or brighter than 17.0 are not included in the AAOmega observation schedule. The green dots represent objects with \\visclass=1; the red crosses \\visclass=2, and the pink crosses \\visclass=3. The small red circles at $r_{\\rm fibre}<17$ are probably stars based on a stricter star-galaxy separation criteria (\\S~\\ref{sec:bright-galaxies}). The blue dash-dotted (dashed) line corresponds to 50\\% (30\\%) redshift success rate for objects on or near the line.} \\label{fig:sb-sep} \\end{figure*} In addition to the explicit SB limits given in Eq.~\\ref{eqn:sb-limits}, which we use to reject objects from our science catalogue, we include a restriction on the fibre magnitudes: \\begin{equation} 17.0 < r_{\\rm fibre} < 22.5 \\label{eqn:fibre-limits} \\end{equation} for targets allocated to the AAOmega observation schedule. This is a practical restriction, with a bright limit to avoid significant crosstalk in the spectrograph and a faint limit because the redshift success is very low. Selected fibre bright targets without a known redshift will be observed with a 2m-class telescope, and, in principle, selected fibre faint targets will be observed with an 8m-class telescope. At the bright end, a more restrictive cut on star-galaxy separation is also justified (see later in \\S~\\ref{sec:bright-galaxies}). \\subsection{Visual classification} \\label{sec:vis-class} Sources with, for example, $\\effsb > 23$ have a high probability of being artifacts, deblends of stars, or the outer parts of galaxies. One of us (J.\\ Liske) has written code to facilitate the visual classification of such sources. A \\visclass\\ variable, initially with zero value, could be changed to the following for each source on inspection: \\begin{itemize} \\item [1] possibly a target, \\item [2] not a target (no evidence of galaxy light), \\item [3] not a target (not the main part of a galaxy). \\end{itemize} First, sources with the following flags all equal to zero, \\textsc{edge, blended, child, maybe\\_cr, maybe\\_eghost}, were assumed to be good, essentially isolated, and not included in any testing (\\visclass\\ set to 255). About 50\\% of targets satisfy these criteria. From the remaining objects, sources were selected for visual classification if any of the following conditions applied: $\\effsb > 23$, $r_{\\rm fibre} > 21$, $r_{\\rm fibre} < 17$, \\textsc{mask\\_ic\\_12} $> 0.2$, $r_{\\rm model} < 15.5$, $r_{\\rm petro} < 15.5$, $r_{\\rm fibre} < r_{\\rm model}$, $r_{\\rm fibre} < r_{\\rm petro}$, near UGC galaxy, within $3''$ of another target, Petrosian radius $>10''$. These indicate that the object could be the result of deblending of a large galaxy, artifact or bright star, e.g., diffraction spikes. In addition to the above criteria, other objects were included in the above process. Objects with the same \\textsc{parentid} as an already classified \\visclass\\ = 3 object were selected. (The above selection was not developed in one go and there have been several iterations.) Finally objects, with the same \\textsc{parentid}, that are the brightest and nearest to any object to be tested were included. Objects that could be part of the same galaxy were viewed together where possible. One had to be certain to classify objects as 3 only if the main part was identified as a target. The above selection produced a sample of about 12\\,500 objects for visual classification, by six observers. Every selected object was classified by three different observers. Of the selected potential main-survey targets (Fig.~\\ref{fig:sb-sep}), \\visclass=1 was set in 92\\% of cases, \\visclass=2 in 5\\% of cases, and \\visclass=3 in 3\\% of cases, based on agreement between two or all three classifiers, 9\\% and 90\\% of cases, respectively. Some of the ambiguous cases were double checked, and a single-observer classification was selected in 1\\% of cases. Objects with values of 2 or 3 were removed from the schedule of AAT observations, i.e., targets must satisfy \\begin{equation} \\mbox{\\visclass} \\neq 2 \\mbox{~~\\textsc{and}~~} \\mbox{\\visclass} \\neq 3 \\: . \\label{eqn:vis-class} \\end{equation} In addition, the $\\visclass=3$ objects can be used to improve the photometry of some large galaxies by coadding in the flux of the galaxy parts (or the `parent' photometry can be used). \\subsection{Number of targets} \\label{sec:number-targs} The total number of objects that are within the GAMA regions (\\S~\\ref{sec:sdss}), main-survey magnitude limits (Eq.~\\ref{eqn:mag-limits}) and $\\dsg > 0.05$, is 143\\,728. Applying the stricter star-galaxy separation (Eq.~\\ref{eqn:star-gal-summary}) reduces the sample to 132\\,073. Removing objects by masking (Eq.~\\ref{eqn:masking}), the SB limits (Eq.~\\ref{eqn:sb-limits}) and visual checking (Eq.~\\ref{eqn:vis-class}), reduces the sample to 120\\,038. Of these, 825 were not included in the AAOmega observation schedule because they do not satisfy the fibre magnitude limits (Eq.~\\ref{eqn:fibre-limits}). A more restictive star-galaxy separation can be applied for brighter targets (discussed later and given in Eq.~\\ref{eqn:star-gal-special}) that reduces the sample to {\\bf 119\\,852}. This is considered to be the main-survey sample. Note these numbers apply to AAT observations in 2009, the numbers may change slightly with addition of complete $J$-$K$ UKIDSS. Separating the main survey into $r$, $z$ and $K$ limited samples, the numbers are 114\\,520, 61\\,418 and 57\\,657, respectively. \\newtext{For $r_{\\rm petro}$ selected samples to 19.0, 19.4 and 19.8, the sample sizes are 60\\,407, 96\\,386 and 150\\,810, respectively. The latter is the $r$-selected main survey plus F2 additional targets, which are described in the following section.} \\subsection{Additional targets} \\label{sec:filler-targets} In order to assess the spectro-photometry of the AAOmega spectra, three or four stars, classified as \\textsc{redden\\_std} or \\textsc{spectrophoto\\_std} by SDSS, were observed in each configuration. These also had a bright fibre magnitude limit of 17 as per the main-survey targets. The aim is to obtain high completeness (99\\%), at least in terms of spectra obtained and ideally in terms of confirmed redshifts, for the main survey. This is set to reduce systematic uncertainties in GAMA's position dependent science cases, and is possible because a given patch of sky is potentially observed by $\\sim5$--10 2dF tiles depending on the local density of targets (see \\citealt{robotham09} for a description of the tiling strategy). Given this requirement, targeting becomes increasingly inefficient as the survey progresses (fewer targets without a redshift per tile). Filler targets were introduced to provide useful redshifts outside the main survey, and thus, maximise fibre usage. These have no high-level requirement on completeness. The filler selections are given by: (F1) objects with detection in the Faint Images of the Radio Sky at Twenty-cm (FIRST) survey and matched to SDSS with $i_{\\rm model} < 20.5$ including unresolved sources; (F2) $19.4 < r_{\\rm petro} < 19.8$ galaxy targets in G09 and G15, aiming for equal depth with G12; and (F3) $g_{\\rm model}<20.6$ or $r_{\\rm model}<19.8$ or $i_{\\rm model}<19.4$ in G12, investigating variation in magnitude-type and wavelength on selection. In total, there are about 50\\,000 filler targets. \\label{sec:summary} The GAMA survey is designed to be a highly complete redshift survey with a target density several times that of SDSS. The survey covers three $48\\,\\sqdeg$ regions near the celestial equator centred on 9\\,h, 12\\,h and 14.5\\,h (Fig.~\\ref{fig:sdss-stripes}). The input catalogue is drawn from the SDSS and UKIDSS. The main-survey limits are $r_{\\rm petro}<19.4$, $z_{\\rm model}<18.2$ and $K_{\\rm AB,auto}<17.6$ ($K<15.7$) across all the regions, and $r_{\\rm petro}<19.8$ over the G12 region (Eq.~\\ref{eqn:mag-limits}). This corresponds to a main survey of 119\\,852 targets. The near-IR selections have a joint constraint with $r_{\\rm model}<20.5$, which has minimal impact on the use of the near-IR selections (Figs.~\\ref{fig:color-bias} \\&~\\ref{fig:obs-color-z}). The GAMA survey lies between that of the SDSS-MGS $r<17.8$ and VVDS-wide $I_{\\rm AB}<22.5$ magnitude-limited samples in the depth-area plane (Fig.~\\ref{fig:compare-grs}). In terms of $K$-band selection (Figs.~\\ref{fig:ukidss-coverage} \\&~\\ref{fig:K-counts}), GAMA covers an area $\\sim15$ times that of the similar-depth Hawaii+AAO $K<15$ survey. In order to be highly complete at the high-SB end of the galaxy distribution, an intensity profile parameter (Eq.~\\ref{eqn:sdss-star-gal-sep}) and a colour-colour parameter (Eq.~\\ref{eqn:ukidss-sdss-star-gal-sep}) are used jointly for star-galaxy separation. The $\\dsgjk$ parameter makes use of $J-K$ and $g-i$ colours. Either parameter works reasonably well in separating stars and galaxies (Figs.~\\ref{fig:sg-sep-histo}--\\ref{fig:test-auto-jk}). A joint selection (Eq.~\\ref{eqn:star-gal-summary}) increases the completeness while stellar contamination in the sample remains at less than 2\\%. Judging by the joint distribution of confirmed galaxies in these parameters (Fig.~\\ref{fig:results-sg-sep}), the completeness is high because the bivariate density drops significantly prior to the limit of our selection. This is particularly important when considering the size evolution of galaxies (Fig.~\\ref{fig:size-z}). The incompleteness at the low-SB end is significant, both in source detection and redshift success rate, which is about 50\\% at $r_{\\rm fibre}=21.5$ (Fig.~\\ref{fig:sb-sep}). Some improvement over the SDSS MGS is made by visually checking low-SB targets ($\\effsb > 23$), rather than using automatic checks, by increased redshift success rate, and by eventually including further integrations of sources with failed redshifts. The GAMA survey has completed two out of a three-year time allocation for spectroscopy with AAOmega on the AAT. To date, 100\\,012 redshifts have been confirmed for the main survey, including 80\\,944 from AAOmega. Of these, 98.5 per cent are extragalactic. \\newtext{The completeness is 96\\% for $r_{\\rm petro}<19.0$, 74\\% for the fainter $r$-band selection, and $\\sim39$\\% for the remaining near-IR selection (Table~\\ref{tab:gama-main-spec}). The completeness at $r>19$ will be significantly improved in the third year of spectroscopic observations.} We expect that this galaxy redshift survey will form a core of a fundamental database for many studies in extragalactic astronomy." }, "0910/0910.0858.txt": { "abstract": "We present a maximum-likelihood analysis for estimating the angular distribution of power in an anisotropic stochastic gravitational-wave background using ground-based laser interferometers. The standard isotropic and gravitational-wave radiometer searches (optimal for point sources) are recovered as special limiting cases. The angular distribution can be decomposed with respect to {\\em any} set of basis functions on the sky, and the single-baseline, cross-correlation analysis is easily extended to a network of three or more detectors---that is, to multiple baselines. A spherical harmonic decomposition, which provides maximum-likelihood estimates of the multipole moments of the gravitational-wave sky, is described in detail. We also discuss: (i) the covariance matrix of the estimators and its relationship to the detector response of a network of interferometers, (ii) a singular-value decomposition method for regularizing the deconvolution of the detector response from the measured sky map, (iii) the expected increase in sensitivity obtained by including multiple baselines, and (iv) the numerical results of this method when applied to simulated data consisting of both point-like and diffuse sources. Comparisions between this general method and the standard isotropic and radiometer searches are given throughout, to make contact with the existing literature on stochastic background searches. ", "introduction": "\\label{s:intro} Data from the laser interferometric gravitational-wave detectors LIGO \\cite{LIGO, Barish:1999,LIGO_eprint,detector}, Virgo \\cite{Virgo, Bradaschia:1990}, and GEO \\cite{GEO-600, Wilke:2004} are currently being analysed for the presence of gravitational waves from a variety of sources. These include signals from inspiraling and coalescing compact binaries (for example, neutron stars and/or stellar mass black holes)~\\cite{bin-insp-S3S4,lowmass_S5,ringdowns_S4,lowmass_S5B}, continuous gravitational waves from quasi-periodic sources such as pulsars~\\cite{SGR,SGR1806,radiopulsars,allsky}, and bursts of gravitational radiation associated with gamma-ray bursts~\\cite{GRB070201,39GRBs,S4GRB}, core-collapse supernovae, or other violent events~\\cite{S3bursts}. In addition, searches are ongoing for the presence of a background of {\\em stochastic} gravitational radiation of either astrophysical or cosmological origin~\\cite{radiometer,LIGO-ALLEGRO,S4Isotropic}, whose detection might provide insights about the very early universe \\cite{Maggiore:2000}, well before the production of the cosmic microwave background \\cite{Kolb:1999}. Although no direct detections of gravitational waves have been made to date, the most recent data taken are of unprecented sensitivity~\\cite{stoch-S5,SGR,GRB070201}, leading to upper limits on gravitational-wave strengths that are competitive with or surpass those from electromagnetic or particle physics observations. Of particular note is the upper limit on the strength of a gravitational-wave signal from the Crab pulsar \\cite{LIGO-crab}, which is a factor of 1.6 lower than the corresponding limit inferred from electromagnetic pulsar spin-down observations~\\cite{palomba}. Also, the current direct limit on the strength of an isotropic stochastic gravitational-wave background at $\\unit[100]{Hz}$ $\\Omega_{\\text gw}<6.9\\times10^{-6}$~\\cite{stoch-S5} (at 95\\% confidence) has surpassed bounds set by considerations from Big Bang Nucleosynthesis~\\cite{Maggiore:2000} and from the microwave background~\\cite{CMB-limit}. In this paper, we describe an analysis method that estimates the angular distribution of power in an {\\em anisotropic} stochastic gravitational-wave background. This method includes both the standard isotropic \\cite{S1Isotropic,S3Isotropic,S4Isotropic} and gravitational-wave radiometer searches \\cite{mitra-et-al,radiometer} (optimal for point sources) as special limiting cases. (For our purposes {\\em anisotropic} is taken to mean {\\em not necessarily isotropic}.) Similar to the radiometer technique, the method presented here looks for modulations in the gravitational-wave signal induced by the Earth's rotational motion relative to an anisotropic background. The method provides maximum-likelihood estimates of the angular distribution of gravitational-wave power ${\\cal P}(\\hat\\Omega)=\\sum_\\alpha {\\cal P}_\\alpha{\\bf e}_\\alpha(\\hat\\Omega)$, decomposed with respect to some set of basis functions on the sky. By choosing a pixel basis %${\\cal P}_{\\hat \\Omega}={\\cal P}(\\hat\\Omega)$, ${\\bf e}_{\\hat\\Omega'}(\\hat\\Omega)=\\delta(\\hat\\Omega,\\hat\\Omega')$, we recover the results of the radiometer method discussed in~\\cite{mitra-et-al,radiometer}. By choosing the spherical harmonics basis $Y_{lm}(\\hat\\Omega)$ defined with respect to the Earth's rotational axis, we obtain maximum-likelihood estimates of the {\\em multipole moments} ${\\cal P}_{lm}$ of the gravitational-wave sky. This basis is particularly convenient as the standard isotropic analysis corresponds to simply restricting attention to the monopole moment ${\\cal P}_{00}$, while the point-source radiometer results are well-approximated by choosing a sufficiently large value of $l_{\\rm max}$ ($l_{\\rm max}\\sim30$), appropriate for the diffraction-limited beam pattern at $f\\sim 1$~kHz. In addition, the use of spherical harmonics simplifies the problem of removing the `smearing' effects of the beam pattern from the measured sky map (that is, deconvolution of the {\\em dirty map}), given the smaller number of elements and symmetries of the beam pattern matrix with respect to the $lm$ indices. The problem of deconvolving a cross-correlated gravitational-wave signal from the interferometers' beam pattern in the spherical harmonic basis has been described in~\\cite{Cornish:2001hg}. We address this problem in detail in section~\\ref{s:data_analysis}. We further note that the spherical harmonic basis is useful for the efficient analysis of cross-correlated data in a variety of applications including searches for transient gravitational-wave sources~\\cite{kipp}. Regardless of basis, the method described here is easily extended to work with a network or three or more detectors with uncorrelated detector noise, by simply adding the individual baseline beam patterns and dirty maps before deconvolution. A multi-baseline analysis improves the overall sensitivity of the search by reducing the variances of the individual estimators, and provides a natural way of regularising the deconvolution of the dirty map; the beam pattern matrix has fewer null (or nearly null) directions for multiple baselines and is thus more stable during inversion. The structure of the rest of the paper is the following: In section~\\ref{s:sgwb}, we briefly review the statistical properties of an anisotropic background, and show how a generalized overlap reduction function arises in a cross-correlation search for such a background. In section~\\ref{s:MLestimation} we derive the optimal estimators of the angular distribution of the gravitational-wave power, starting from the likelihood function for cross-correlated data. We explicitly construct the beam pattern matrix, and discuss its relation to the covariance matrix of the estimated ${\\cal P}_{\\alpha}$. Section \\ref{s:data_analysis} describes details of the data analysis implementation and issues related to deconvolution and regularisation. %Section \\ref{s:multiple_baselines} It also briefly describes how to extend %the results of the previous sections the analysis to a network of three or more detectors, and the expected increase in sensitivity from using multiple baselines. In section~\\ref{s:simulations} we present numerical results of the method applied to simulated data. We consider both point-like and diffuse-source injections, and compare the extracted and injected sky-maps. Finally, in section~\\ref{s:summary} we summarize our results. We also include three appendices: Appendices ~\\ref{s:Ylms} and \\ref{s:identities} contain definitions of the spherical harmonics and some useful identities relating different multipole moments, beam pattern matrix components, etc.; Appendix \\ref{s:DetStat} defines a related detection statistic that assumes a particular distribution of (normalized) angular distribution functions on the sky. ", "conclusions": "\\label{s:summary} We have presented here a maximum-likelihood analysis method for estimating the angular distribution of power in an anisotropic stochastic gravitational-wave background. The basic idea was to cross-correlate data from a network of two or more gravitational-wave detectors, exploiting time-of-arrival differences and the diurnal modulation due to the Earth's rotation. We derived maximum-likelihood estimators for the angular distribution of gravitational-wave power ${\\cal P}(\\hat\\Omega)=\\sum_\\alpha {\\cal P}_\\alpha \\mathbf{e}_\\alpha(\\hat\\Omega)$, decomposed with respect to {\\em any} set of basis functions on the sky. We derived an expression for the beam pattern matrix $\\Gamma_{\\alpha\\beta}$ and discussed its relationship to the covariance matrix %$(\\Gamma^{-1})_{\\alpha\\beta}$ of the maximum-likelihood estimators $\\hat{\\cal P}_\\alpha$. We described how singular value decomposition can be used to regularize the inverse of $\\Gamma_{\\alpha\\beta}$, which was needed to remove the smearing effects of the beam pattern matrix on the measured (`dirty') sky maps $X_\\alpha$. We also explained how the single-baseline (two-detector) cross-correlation analysis can be extended to a network of three or more detectors, thereby increasing our sensitivity to detecting a signal. In this paper, we focused attention on a decomposition with respect to a basis of spherical harmonics $Y_{lm}(\\hat\\Omega)$, for which the maximum-likelihood estimators $\\hat{\\cal P}_{lm}$ represent the multipole moments of the gravitational-wave sky, and for which the standard isotropic %\\cite{S1Isotropic,S3Isotropic,S4Isotropic} and radiometer %\\cite{mitra-et-al, radiometer} searches are recovered as special limiting cases. Finally, we illustrated all these general results by analysing simulated data containing injected stochastic gravitational-wave backgrounds having different angular power distributions." }, "0910/0910.1779_arXiv.txt": { "abstract": "The radio emitting X-ray binary GRS\\,1915+105 shows a wide variety of X-ray and radio states. We present a decade of monitoring observations, with the RXTE-ASM and the Ryle Telescope, in conjunction with high-resolution radio observations using MERLIN and the VLBA. Linear polarisation at 1.4 and 1.6~GHz has been spatially resolved in the radio jets, on a scale of $\\sim150$~mas and at flux densities of a few mJy. Depolarisation of the core occurs during radio flaring, associated with the ejection of relativistic knots of emission. We have identified the ejection at four epochs of X-ray flaring. Assuming no deceleration, proper motions of 16.5 to 27~mas per day have been observed, supporting the hypothesis of a varying angle to the line-of-sight per ejection, perhaps in a precessing jet. ", "introduction": "GRS\\,1915+105 has proved to be one of the best laboratories for the study of relativistic physical environments, due to its high-brightness, periodic outbursts of superluminal ejecta and relative proximity. Since the first detection in August 1992 with the WATCH instrument on board the X-ray telescope GRANAT~\\citep{1992IAUC.5590....2C}, GRS\\,1915+105 has demonstrated some of the most spectacular high-energy physics within the Galaxy. Shortly after its discovery, a radio counterpart was discovered by \\cite{1993IAUC.5773....2M} and later an infrared component was also identified \\citep{1993IAUC.5830....1M}. Its radio light-curve shows a highly variable and complex structure, with spatially resolved features that can be followed over many days. A clear correlation between the X-ray, infrared and radio emission was quickly established by~\\cite{1994Natur.371...46M}, with rapid time variability in each band. Radio flaring was found to correspond to a rapid change in the hard X-rays and possible production of high-energy gamma rays. In March 1994, 20~cm VLA observations of GRS\\,1915+105 detected the first superluminal motion of a Galactic source \\citep{1994Natur.371...46M}. This major breakthrough provided the direct evidence of relativistic jets and an extreme physical environment within the Galaxy. The name ``microquasar'' was coined due to their obvious similarities with their extra-galactic counterparts, quasars. The launch of the Rossi X-ray Timing Explorer (RXTE) satellite, in December 1995, signified the start of a long-term monitoring campaign of X-ray binaries. The All Sky Monitor (ASM) on board the RXTE has taken daily observations of GRS\\,1915+105 since its launch. \\cite{1996ApJ...473L.107G} detected unusual X-ray variability on time scales of under one second to days. The RXTE-ASM lightcurve was found to be both highly complex and structured, due to instabilities in the accretion disc. Detailed observations with the RXTE's Proportional Counter Array (PCA) instrument (\\citealt{1997ApJ...477L..41C}; \\citealt{1997ApJ...488L.109B}) identified different spectral states, including a low-hard state dominated by a power-law and a high-soft state with a strong disc-blackbody component. A transition between such states is believed to be associated with the ejection of superluminal plasmons \\citep{1999MNRAS.304..865F}. \\begin{table*} \\begin{center}\\begin{tabular}{|>{\\columncolor[rgb]{0.8,0.8,0.8}}l|ccccccc|} \\hline \\rowcolor[rgb]{0.8,0.8,0.8} Epoch (MJD) & Frequency/Array & $I_{\\rm{total}}$ & $LP_{\\rm{total}}$ & Fraction & Noise & Spectral$^{\\ddagger}$ & Resolved? \\\\ \\rowcolor[rgb]{0.8,0.8,0.8} & (GHz) & (mJy) & (mJy) & (\\%) & ($\\mu$Jy/beam) & Index & \\\\ \\hline $^{\\star}$1997 October 31 (50752) -core & 2.27 / VLBA & $105\\pm2$ & $\\le0.1^{\\dagger}$ & $\\le0.5^{\\dagger}$ & 430 & \\multirow{4}{*}{-} & \\multirow{4}{*}{YES} \\\\ $^{\\star}$1997 October 31 (50752) -jet & 2.27 / VLBA & $226\\pm2$ & $5.6\\pm0.3$ & 2.5 & 430 & & \\\\ $^{\\star}$1997 October 31 (50752) -core & 8.3 / VLBA & $33\\pm1$ & $\\le0.1^{\\dagger}$& $\\le0.5^{\\dagger}$ & 460 & &\\\\ $^{\\star}$1997 October 31 (50752) -jet & 8.3 / VLBA & $65\\pm3$ & $4.7\\pm0.3$& 7.2 & 460 & &\\\\ \\hline 1998 June 9 (50973) & 1.6 / MERLIN & $145.1\\pm0.4$ & $9.5\\pm0.3$& $6.9\\pm0.2$ & 300 & -0.3 & NO\\\\ \\hline 1999 April 2 (51270) & 1.6 / MERLIN & 238.4$\\pm0.3$ & $3.5\\pm0.3$& $1.5\\pm0.2$ & 530 & -0.6& NO\\\\ \\hline 1999 November 18 (51500) & 1.6 / MERLIN & 221.3$\\pm0.8$ & $6.9\\pm0.3$ & $3.2\\pm0.2$ & 860 & -0.4& NO\\\\ \\hline 1999 November 22 (51504) & 1.6 / MERLIN & 37.1$\\pm0.7$ & $\\le0.1^{\\dagger}$ & $\\le0.4^{\\dagger}$ & 300 & -0.2& NO\\\\ \\hline 1999 December 28 (51540) & 1.6 / MERLIN & 89.1$\\pm0.7$ & $12.7\\pm1.6$ & $14.1\\pm0.2$ & 400 & -0.4& NO\\\\ \\hline 2003 March 6 (52704) & 1.6 / MERLIN & $31.3\\pm0.3$ & $1.6\\pm0.3$ & $5.2\\pm1.0$ & 240 & -0.4& NO\\\\ \\hline 2003 March 25 (52723) & 1.6 / MERLIN & $125.8\\pm0.1$ & $7.1\\pm0.2$ & $5.6\\pm0.2$ & 140 & 0.2& NO\\\\ \\hline 2003 April 18 (52747) -core & 1.6 / MERLIN & $135.4\\pm0.4$ & $3.5\\pm0.2$ & $2.6\\pm0.2$ & 270 & \\multirow{2}{*}{0.0}& \\multirow{2}{*}{YES}\\\\ 2003 April 18 (52747) -jet & 1.6 / MERLIN & $20.0\\pm0.3$ & $1.3\\pm0.1$ & $6.5\\pm0.5$ & 270 & &\\\\ \\hline 2003 May 09 (52768) & 1.6 / MERLIN & $43.0\\pm0.1$ & $2.7\\pm0.1$ & $6.3\\pm0.2$ & 160 & 0.2&NO\\\\ \\hline 2003 June 15 (52805) -core & 1.6 / MERLIN & $41.4\\pm0.1$ & $\\le0.1^{\\dagger}$ & $\\le0.2^{\\dagger}$ & 390 & \\multirow{2}{*}{-0.4}& \\multirow{2}{*}{YES}\\\\ 2003 June 15 (52805) -jet & 1.6 / MERLIN & $67.5\\pm0.1$ & $7.6\\pm0.4$ & $11.9\\pm0.6$ & 390 & & \\\\ \\hline 2006 December 24 (54093) & 1.4 / MERLIN & $145.0\\pm0.5$ & $7.7\\pm0.6$ & $5.3\\pm0.4$ & 380 & \\multirow{2}{*}{-0.1}& \\multirow{2}{*}{NO}\\\\ 2006 December 24 (54093) & 1.6 / MERLIN & $156.9\\pm1.2$ & $11.5\\pm0.5$ & $7.3\\pm0.5$ & 380 && \\\\ \\hline 2006 December 27 (54096) & 1.4 / MERLIN & $20.7\\pm0.5$ & $\\le0.1^{\\dagger}$ & $\\le0.5^{\\dagger}$ & 230 & \\multirow{2}{*}{-0.4} & \\multirow{2}{*}{NO}\\\\ 2006 December 27 (54096) & 1.6 / MERLIN & $27.0\\pm0.5$ & $\\le0.1^{\\dagger}$ & $\\le0.5^{\\dagger}$ & 230 & &\\\\ \\hline 2006 December 28 (54097) & 1.4 / MERLIN & $25.8\\pm0.4$ & $\\le0.1^{\\dagger}$ & $\\le0.4^{\\dagger}$ & 210 & \\multirow{2}{*}{-}& \\multirow{2}{*}{NO}\\\\ 2006 December 28 (54097) & 1.6 / MERLIN & $26.2\\pm0.4$ & $\\le0.1^{\\dagger}$ & $\\le0.4^{\\dagger}$ & 210 & &\\\\ \\hline 2007 January 04 (54104) & 1.4 / MERLIN & $21.0\\pm0.5$ & $\\le0.1^{\\dagger}$ & $\\le0.5^{\\dagger}$ & 260 & \\multirow{2}{*}{-0.3}& \\multirow{2}{*}{NO}\\\\ 2007 January 04 (54104) & 1.6 / MERLIN & $22.2\\pm0.5$ & $\\le0.1^{\\dagger}$ & $\\le0.5^{\\dagger}$ & 260 & &\\\\ \\hline \\end{tabular}\\end{center} \\caption{\\label{table:journal}Journal of observations and results: Date of observation, observed frequency and array, total flux, total linearly polarised flux, fraction of total flux that is linearly polarised, systematic noise, spectral index and detection of resolved structure. [Notes: $^{\\star}$ The observations were first published by \\citealt{2000ApJ...543..373D}; $^{\\dagger}$ Estimates are based on a upper limit of detection; $^{\\ddagger}$ The spectral index is derived from the total MERLIN flux (i.e. core + jet) and the 15 GHz RT data, where $S \\propto \\nu^\\alpha$.]} \\end{table*} Radio polarisation was first detected with MERLIN in GRS\\,1915+105 by \\cite{1999MNRAS.304..865F} in the first four epochs (i.e. first four days) of the October 1997 flare at a frequency of 5~GHz. A clear asymmetry in the location of the polarised emission was found, showing the strongest detection of linearly polarised emission within the approaching jet, weaker in the receding jet and no detection ($<3\\%$ of the total flux density) within the core. The polarisation position angle (PA) was found to rotate by at least $75^\\circ$ over the four days, leading to the suggestion that the changes in polarisation could be due to Faraday effects (i.e. changes in the Faraday depth as the plasmons move away from the core), implying a change in rotation measure of $\\Delta RM>300$~rad~m$^{-2}$. This paper describes observations taken over a decade with MERLIN at 18 and 21~cm. The observations were either in conjunction with INTEGRAL observations as part of the Galactic Plane Survey~\\citep{2004ApJ...607L..33B,2004ESASP.552..321F}, or were triggered by flare events, found by other radio telescopes. Polarisation behaviour, structural variations and the relationship of activity in the radio regime to the X-ray behaviour were investigated. ", "conclusions": "GRS\\,1915+105 has been simultaneously observed in the radio and X-rays in various different states over a ten year period. This comparative study between the X-ray and radio variability shows, for the first time, both long term ($\\sim$~weeks/months) and short lived (intra-day) variations. Daily monitoring shows periods of extended activity and X-ray spectral states that are related to the ejection of material. Radio activity over this period has shown to originate from the central 150 mas core, unless a short X-ray flare is observed that marks the ejection of a superluminal plasmon. The proper motion of ejecta has been calculated by identifying the time of ejection and exhibit no deceleration. Measured velocities were in agreement with previous observations, and show a significant variation in velocity. Polarisation measurements show strong linear polarisation variation. Depolarisation in the core of the XRB is found during the ejection of a plasmon and the multi-wavelength observations show an increase of linear polarisation with frequency, suggesting internal Faraday rotation effects. Observing GRS\\,1915+105 over a long period of time has shown the relativistic out-flow in different forms. If the accretion conditions permit, collimated and steady jets persist over many weeks. Different apparent velocities of superluminal knots suggest either a precession of the jet angle or variation in their formation. Finally, the nature of the short-lived radio flares remains unknown; for example, it is difficult to explain why these bright, but short-lived events do not produce extended structure, in contrast to a flare during a state transition." }, "0910/0910.4640_arXiv.txt": { "abstract": "We processed the data about radial velocities and HI linewidths for $1678$ flat edge-on spirals from the Revised Flat Galaxy Catalogue. We obtained the parameters of the multipole components of large-scale velocity field of collective non-Hubble galaxy motion as well as the parameters of the generalized Tully-Fisher relationship in the ``HI line width -- linear diameter'' version. All the calculations were performed independently in the framework of three models, where the multipole decomposition of the galaxy velocity field was limited to a dipole, quadrupole and octopole terms respectively. We showed that both the quadrupole and the octopole components are statistically significant. On the basis of the compiled list of peculiar velocities of $1623$ galaxies we obtained the estimations of cosmological parameters $\\Omega_m$ and $\\sigma_8$. This estimation is obtained in both graphical form and as a constraint of the value $S_8=(\\Omega_m/0.3)^{0.35} \\sigma_8 = 0.91 \\pm 0.05$. ", "introduction": "\\label{s:Introduction} The distribution of matter, including dark matter, in the Universe is inhomogeneous on the scales of about $100\\,Mpc$. This is manifested, for example, in the existence of superclusters and voids. A galaxy, besides the cosmological expansion, is also attracted to the regions with greater density. As a result, the galaxies are involved in a non-Hubble large-scale collective motion on the background of Hubble expansion. In a more sophisticated way this can be considered as a development of initial density and velocity fluctuations in the early Universe due to the gravitational instability. Investigation of such a motion is important at least for two reasons. First of all, it allows to plot the distribution of matter, including dark matter, in the surrounding region of the Universe, which is the main goal of cosmography, and to compare this distribution with the distribution of luminous matter. The second reason is that the parameters of this motion are linked with certain cosmological parameters, so we can obtain new independent estimations of these parameters. Of course, the accuracy of such estimation will not be very high, but the agreement of different estimations of cosmological parameters can support the correctness of the standard cosmological model. In 1989 \\citeauthor{ref:K89} proposed to use flat edge-on spiral galaxies as a tool for studying their large-scale collective motion. They are good in this role for the following reasons: \\begin{enumerate} \\item{The linear diameter is strongly correlated with the HI linewidth for thin bulgeless galaxies. This allows to determine distances without photometric data.} \\item{Flat galaxies can be easily identified by their axes ratio.} \\item{Flat galaxies have a nearly 100 per cent HI detection rate.} \\item{Galaxies in clusters are usually not flat due to interaction with neighbours. This means that flat galaxies avoid clustering and do not interact with the intergalactic gas in clusters.} \\item{Peculiar velocities of isolated flat galaxies are not perturbed by neighbours.} \\end{enumerate} To implement this idea the Flat Galaxy Catalogue \\citep[FGC,][]{ref:FGC} was created. It contained data about $N=4455$ galaxies, which satisfied the conditions $a_b/b_b\\ge7$ and $a_b>0.6'$. Here $a_b$ and $b_b$ are the major and minor axial diameters directly measured on POSS-I and ESO/SERC plates. In accordance with the original photographic material, the Catalogue consists of two parts: FGC ($N=2573$) and its southern extension, FGCE ($N=1882$). The first part is based on the POSS-I and covers the sky region with declinations between $-20\\deg$ and $+90\\deg$. The second one is based on the ESO/SERC and covers the rest of the sky area up to $\\delta=-90\\deg$. After thorough studies of the catalogue's properties, both these parts were joined. The angular diameters from the FGCE were converted to the POSS-I system of the FGC, which appeared to be close to the $a_{25}$ system. This system, where galaxy size is taken at $B=25\\,mag/\\square''$ isophotal level, was used, in particular, by \\citet{ref:deVac}. Some FGCE galaxies, which did not satisfy the condition $a>0.6'$ after conversion, were removed from the resulting Revised Flat Galaxy Catalogue \\citep[RFGC,][]{ref:RFGC}. It contained data about $N=4236$ galaxies including the information on the following parameters: Right Ascension and Declination for the epochs J2000.0 and B1950.0, galactic longitude and latitude, major and minor blue and red diameters in arcminutes in the POSS-I diameter system, morphological type of the spiral galaxies according to the Hubble classification, index of the mean surface brightness (I -- high, IV -- very low) and some other parameters, which are not used in this article. More detailed description of the catalogue can be found in the paper \\citep{ref:RFGC}. The original goal of this catalogue was to estimate the distance to galaxies according to the Tully-Fisher relationship in the ``HI line width -- linear diameter'' version without using their redshifts. The difference between the velocity $V$ derived from the redshift and the Hubble velocity $R=Hr$ corresponding to the distance $r$ estimated by Tully-Fisher relationship is called a peculiar velocity $V_{pec}=cz-Hr$. We can use such a simple form of the Hubble's law because our sample has $z<0.1$. There are some important things to take into account about peculiar velocities. The redshift includes not only the velocity of the galaxy, but also the velocities of our Galaxy, Solar System and the Earth. Thus, to eliminate these factors, all velocities were reduced to the frame, where CMB is isotropic. Naturally, the redshift gives us only the radial component of the velocity and the tangential components cannot be measured. Additionally, Tully-Fisher relationship is statistical and thus has a certain error. Thus, we can only estimate the peculiar velocity for each galaxy, sometimes with a significant error. However, we believe that there is a large-scale velocity field. We consider the individual galaxies as test particles in the velocity field of large-scale collective motion. Thus, having data about the peculiar velocities of a large number of galaxies, we can restore the radial component of the velocity field. Using some additional assumptions, like the potentiality of the flux, it is possible to restore the 3D velocity field. For this reason we need ample samples of peculiar velocities, preferably uniformly covering the celestial sphere. These conditions are satisfied by the RFGC catalogue, which covers the whole celestial sphere in both hemispheres with the natural exclusion of the Milky Way region. However, not all of the RFGC galaxies have data required to estimate the distance to them using the Tully-Fisher relationship. Nevertheless, the sample of galaxies having such data is also quite uniform, as shown on Fig. \\ref{fig:1}. \\begin{figure*}[tb] \\includegraphics[width=\\textwidth]{Parnovsky_fig1.eps} \\caption{Distribution of galaxies over the celestial sphere in galactic coordinates} \\label{fig:1} \\end{figure*} Naturally, RFGC is not the only sample that can be used for this purpose. Possible alternatives include SBF \\citep{ref:SBF97,ref:SBF00}, ENEAR \\citep{ref:ENEAR}, SFI \\citep{ref:SFIa,ref:SFIb}, SFI++ \\citep{ref:SFIpp}, EFAR \\citep{ref:EFAR96,ref:EFAR01}, SMAC \\citep{ref:SMAC99,ref:SMAC04}, 2MFGC \\citep{ref:2MFGC}. However, in this article we use only the RFGC catalogue. To apply the Tully-Fisher relationship the data from the catalogue is not sufficient; we also need to know the HI linewidth $W$ (in this article we take it at $50$ per cent level), or the gas rotation velocity $V_{rot}$ obtained from optical observations. Additionally, we need redshift data. The number of galaxies with such data increases constantly. Since FGC and RFGC were assembled, a number of articles were published dealing with collective motions of RFGC galaxies. Very preliminary results were published by \\citet{ref:K95}. The parameters of the bulk motion were calculated by \\citet{ref:K00}. In the paper \\\\ \\citep{ref:Par01} not only new data were added, but also the models featuring the quadrupole and octopole components of the velocity field were proposed. Also, the generalized Tully-Fisher relationship for RFGC galaxies was finalized. It included data not only about HI linewidth and angular diameter in red and blue bands, but also about the morphological type of the galaxy and its surface brightness index. By that time the authors had information about radial velocities and HI $21\\,cm$ line widths or $V_{rot}$ for $1327$ RFGC galaxies from the different sources listed in the paper \\citep{ref:K00}. From this number, $1271$ galaxies were included in the sample, and the rest of them were considered to be outliers. As a result, the first list of peculiar velocities of RFGC galaxies was prepared by \\citet{ref:list}. Four years later, the number of galaxies with available data increased and reached $1561$ \\citep{ref:ParTug04}; $1493$ of them entered the sample. A new version of the list of peculiar velocities was prepared by \\citet{ref:ParTug05}. This list was the basis for solving the two abovementioned problems, namely mapping the matter density and estimation of cosmological parameters. The distribution of matter density was obtained in the paper \\citep{ref:SharPar} up to $75h^{-1}\\,Mpc$ in the supergalactical plane and 8 more planes. In the same article, the excess masses of attractors in this region were estimated, and the estimation of $\\beta$ parameter was obtained. The estimations of the cosmological parameters $\\Omega_m$ and $\\sigma_8$ were obtained in the paper \\citep{ref:cosm}. The model was expanded taking into account the effects of general theory of relativity in the paper \\citep{ref:ParGayd05}. An important result, obtained in the paper \\\\ \\citep{ref:Par01}, and confirmed by \\\\ \\citet{ref:ParTug04}, was the statistical significance of quadrupole and octopole components (both at more than $99$ per cent confidence level). This result relied on the F-test, which assumes normal distribution of errors \\citep[see][]{ref:Hudson}. However, even the normally distributed errors of angular diameters and HI linewidths and deviations from Tully-Fisher relationship lead to non-Gaussian distribution of peculiar velocities. Thus, the statistical significance of these terms required additional consideration. In the article \\citep{ref:ParPar08} it was shown using Monte-Carlo simulations that the statistical significance of these components appeared to be less than perceived from the F-test, but still large enough to be considered: the quadrupole was statistically significant at $96$ per cent confidence level and the octopole -- at $90..93$ per cent confidence level. In the four years that passed since the last list of peculiar velocities of RFGC galaxies was compiled, not only new data have appeared, but also some old data were remeasured. This led to a necessity of repeating all the steps required to obtain the new list. This means that not only new data have to be added, but the process of sample selection must be repeated from the very beginning. As it will be shown further, some results appeared to be notably different from the ones obtained with the previous sample. In this paper we present the parameters of the large-scale collective velocity field in the framework of the multipole model as well as the estimations of some cosmological parameters. The distribution of matter density will be discussed in a separate article. ", "conclusions": "\\label{s:Conclusion} We prepared a new increased and largely revised sample of RFGC galaxies with data about redshifts and HI linewidths. It allowed us to improve the estimations of parameters of the radial velocity field of large-scale collective motions of galaxies using 3 models of its multipole structure. In comparison with the previous versions the main features remained intact, but separate details, for example the apex of the dipole component of the bulk motion in the D-model, significantly changed. As before, the statistical significance according to F-test of both the quadrupole and the octopole components is well above $99.5$ per cent. The exact parameters are given in the article for subsamples with maximum distances $100h^{-1}\\,Mpc$ and $80h^{-1}\\,Mpc$. The obtained velocity of the bulk motion is in agreement with the $\\Lambda$CDM model expectation of $\\sim 200\\,km\\,s^{-1}$. The value of the bulk motion velocity attracted additional attention after the recent paper of \\citet{ref:WFH09} who obtained this value as $407 \\pm 81\\,km\\,s^{-1}$ at the same scale $100h^{-1}\\,Mpc$ as in present article. In this connection some authors immediately started speculating about the challenge to the $\\Lambda$CDM model and the necessity for inclusion of non-gravitational forces \\citep{ref:AWW09}. For $1623$ galaxies we compiled a list of peculiar velocities which will be published in the nearest future. We also intend to use it to determine the distribution of the matter density (including dark matter) at the scales $75h^{-1}\\,Mpc$ using POTENT method. This list was also used to estimate the cosmological parameters $\\Omega_m$ and $\\sigma_8$. The obtained constraint is given in graphical form on Fig. \\ref{fig:6}. The best numerical constraint is given by a combination $S_8=(\\Omega_m/0.3)^{0.35} \\sigma_8=0.91\\pm 0.05$." }, "0910/0910.3060_arXiv.txt": { "abstract": "{Solar-like oscillations have now been observed in several stars, thanks to ground-based spectroscopic observations and space-borne photometry. CoRoT, which has been in orbit since December 2006, has observed the star HD49933 twice. The oscillation spectrum of this star has proven difficult to interpret. } {Thanks to a new timeseries provided by CoRoT, we aim to provide a robust description of the oscillations in HD49933, i.e., to identify the degrees of the observed modes, and to measure mode frequencies, widths, amplitudes and the average rotational splitting.} {Several methods were used to model the Fourier spectrum: Maximum Likelihood Estimators and Bayesian analysis using Markov Chain Monte-Carlo techniques.} {The different methods yield consistent result, and allow us to make a robust identification of the modes and to extract precise mode parameters. Only the rotational splitting remains difficult to estimate precisely, but is clearly relatively large (several $\\mu$Hz in size).} {} ", "introduction": "\\label{sec:intro} Stars are now the objects of seismic studies after decades of similar studies for the Sun, thanks to the advent of space-borne photometric observations (e.g., MOST, CoRoT and Kepler) and extremely precise ground-based spectroscopic observations \\citep[for a complete review, see for e.g.][]{Aerts2008SoPH}. This applies in particular to stars presenting solar-like p modes (acoustic oscillations stochastically excited by convection), with CoRoT observations of such stars showing clearly individual peaks in the Fourier spectra \\citep[e.g.][]{michel08}. Among these stars, HD49933 has already been the target of asteroseismic campaigns. It was first observed spectroscopically from the ground for 10 nights \\citep{mosser05}. In 2007, a first photometric time series of 60 days was collected by CoRoT, followed by a new long run of 137 days in 2008. HD49933 is a F5 main sequence star with an apparent visual magnitude $m_V = 5.77$. It is hotter than the Sun \\citep[$T_{\\rm eff}$=6780\\,K or $T_{\\rm eff}$=6500\\,K,][]{bruntt08,bruntt09,ryabchi09} with an estimated mass of $\\sim 1.2M_{\\odot}$ \\citep{mosser05} and an estimated radius of $1.34 \\pm 0.06\\,R_{\\odot}$ \\citep{thevenin06}. The surface rotation ($v\\sin i$) was determined to be around 10\\,km\\,s$^{-1}$ \\citep{mosser05,solano05}. The surface rotation period has also been measured at $\\sim$\\,3.4 days, using the 60-day CoRoT timeseries \\citep{appourchaux08,deheuvels2008}, from the signatures of photospheric transiting active regions (e.g., spots) which give rise to a clear peak in the very low-frequency part of the Fourier spectrum. The seismic interpretation of HD49933 has proven to be very difficult. \\citet{mosser05} could not isolate individual p modes in the Fourier spectrum of observed line-of-sight velocities but were able to find a regular pattern in the spectrum, which is the signature of the large frequency separation between modes of same degree $l$ (but increasing radial order $n$). The first CoRoT time-series was analysed by \\citet{appourchaux08}, and these data clearly show individual p-mode peaks in the Fourier spectrum. However, the peaks show large widths, making the interpretation less than straightforward: a given peak could be interpreted as being a closely spaced pair of $l$=0 and $l$=2 modes, or a single (but rotationally split) $l$=1 mode. Based on the modeling of the spectrum using a Maximum Likelihood Estimator fitting method, \\citet{appourchaux08} chose one of these two possible interpretations (hereafter called model A) based on the highest {\\em likelihood} of each model. This first time-series was the object of other studies. \\cite{appourchaux2009a} put into perspective the results of \\citet{appourchaux08}, showing that the likelihood ratio test does not give the probability of the hypothesis given the data, but only the significance of the data given the hypothesis. \\cite{benomar09}, who applied a Bayesian analysis to the same time series, could not definitely favour one interpretation (model A) over the alternate (model B), based on the {\\em whole} probability distribution of each model. \\citet{gruberbauer2009a}, using a Bayesian approach too, also consider the identification ambiguous. \\citet{gaulme2009} used a simpler Bayesian approach (Maximum A Posteriori, or MAP, approach). The most probable model they found corresponds to the same identification as \\citet{appourchaux08}. More recently, \\citet{mosser09b} proposed an empirical method to determine the identification of the modes. Its application considers the model B as the more likely when using the two datasets used here. It should be noted that the case of HD49933 is quite different from the solar case: solar modes are very narrow in comparison. Their widths (1\\,$\\mu$Hz for the modes with the highest amplitudes) are much smaller than the small-frequency separations between $l$=0 modes of order $n$ and the neighbouring $l$=2 modes of order $n$-1 (being typically around 10\\,$\\mu$Hz for the Sun). Moreover, the star inclination angle, if small, tends to attenuate the visibility of mode components with azimutal order $m\\ne0$, making the mode identification even more difficult. Here, we use the new long-run CoRoT observations of HD49933, together with the original shorter run timeseries (Fig.\\,\\ref{fig:LC}), in order to properly describe the acoustic oscillations of the star clearly visible in the Fourier spectrum (Fig.\\,\\ref{fig:spec}, Fig.\\,\\ref{fig:diag_ech0} and Fig.\\,\\ref{fig:zoomspec}). \\begin{figure*} \\begin{center} \\includegraphics[angle=90,width=18.4cm,height=6.8cm]{13111A.ps} \\end{center} \\caption{Mean power spectrum using 3 time series of 60 days (solid grey line), and the fitted model (dashed line).} \\label{fig:spec} \\end{figure*} ", "conclusions": "\\label{sec:conc} The two CoRoT observation runs on the star HD49933 -- the first 60-day run, and the more recent longer run -- have now provided enough data to resolve the identification of modes in the oscillation spectrum at a very high confidence level. This identification relied on several independent analyses. The new data have also allowed us to improve the precision in the mode parameters, with fractional improvements being in the range from 40\\,\\% to 70\\,\\% depending on the parameter. It is now possible to determine precise mode frequencies for $l$=0 and 1 modes. The $l$=2 mode parameters are more difficult to estimate because of the overlap with the stronger, neighbouring $l$=0 modes. The widths and amplitudes of the modes are well determined, as is the inclination angle of the star. The rotational frequency splitting remains the only mode parameter that is poorly constrained. However, it is clearly much higher than the solar value, and similar or larger in size to the inverse of the surface rotation period. In summary, this new information on the acoustic oscillations of HD49933 now opens the possibility for detailed seismic modeling of the star." }, "0910/0910.0580_arXiv.txt": { "abstract": "Double-peaked \\oiii\\, profiles in active galactic nuclei (AGNs) may provide evidence for the existence of dual AGNs, but a good diagnostic for selecting them is currently lacking. Starting from $\\sim$ 7000 active galaxies in SDSS DR7, we assemble a sample of 87 type 2 AGNs with double-peaked \\oiii\\, profiles. The nuclear obscuration in the type 2 AGNs allows us to determine redshifts of host galaxies through stellar absorption lines. We typically find that one peak is redshifted and another is blueshifted relative to the host galaxy. We find a strong correlation between the ratios of the shifts and the double peak fluxes. The correlation can be naturally explained by the Keplerian relation predicted by models of co-rotating dual AGNs. The current sample statistically favors that most of the \\oiii\\ double-peaked sources are dual AGNs and disfavors other explanations, such as rotating disk and outflows. These dual AGNs have a separation distance at $\\sim 1$ kpc scale, showing an intermediate phase of merging systems. The appearance of dual AGNs is about $\\sim 10^{-2}$, impacting on the current observational deficit of binary supermassive black holes with a probability of $\\sim 10^{-4}$ (Boroson \\& Lauer). ", "introduction": "Supermassive black hole binaries and dual AGNs are the natural result of the current $\\Lambda$CDM cosmological paradigm, in which galaxies grow hierarchically through minor or major mergers. However, evidence for such systems is elusive and confirmed cases are rare despite circumstantial evidence. Examples can be found in Kochanek et al. (1999), Junkkarinen et al. (2001), Ballo et al. (2004), Maness et al. (2004), Guainazzi et al. (2005), Rodriguez et al. (2006), Hudson et al. (2006), Barth et al. (2008). The best-known examples of dual AGNs are found in NGC 6240 (Komossa et al. 2003), EGSD2 J142033+525917 (Gerke et al. 2007), EGSD2 J141550+520929 (Comerford et al. 2009a), the $z=0.36$ galaxy COSMOS J100043+020637 (Comerford et al. 2009b). Ideally, analysis of a large statistical sample should be undertaken. The narrow emission lines in AGNs are generally thought to be produced by clouds in the narrow line regions at a scale of $\\sim 1$ kpc, where they are photoionized by the central energy source (e.g. Netzer et al. 2006). It is common that the \\oiii$\\lambda$5007 line has an asymmetric blue wing (Boroson 2005, Komossa et al. 2008), which could be caused by winds and outflows. On the other hand, about 1\\% AGNs have double-peaked \\oiii\\, profiles. The origin of the double-peaked emission lines remains unclear, but they have been interpreted as evidence for bi-polar outflows in the early 1980s (Heckman et al. 1981, 1984; Whittle 1985a,b,c) or disk-shaped narrow line regions (Greene et al. 2005). Evidence based on objects mentioned above suggests dual AGNs as an alternative explanation of the double-peaked profiles. However, more robust observational evidence is needed. In this Letter, we select type 2 AGNs from SDSS DR7. The obscuration of the nucleus in these sources allows us to determine the redshifts of the host galaxies through the stellar absorption lines and investigate the properties of the double peaks. We find a strong correlation between the ratios of the shifts and the double peak fluxes. This is most naturally explained by the Keplerian relation of co-rotating dual AGNs. ", "conclusions": "We find a striking correlation between the ratios of \\oiii shifts and the double peak fluxes in a sample of 87 AGNs from the SDSS. This correlation strongly favors the explanation of dual AGNs in light of the Kepler diagram and does not favor a disk and outflow origin. The results also show that dual AGNs at 1 kpc scale are common, strongly impacting on the observational deficit of binary SMBHs with a probability of $\\sim 10^{-4}$ as shown by Boroson \\& Lauer (2009). The present correlation in the Kepler diagram and the dynamical criterion offer an efficient way to select targets (e.g. full version of Table 1) for future image detections in optical band and X-rays. Finally, redshifts of host galaxies are a vital parameter in determination of the position of sources in the Kepler diagram. Higher resolution spectra are extremely important to diagnose dual AGNs among the weak double-peaked sources. The method presented in this paper is ideal for the study of such sources using high resolution spectra in the future." }, "0910/0910.2911_arXiv.txt": { "abstract": "We present our ground-based CCD observations of the close binary systems DD~Mon and XY~UMa in B, V, R and I bands. The light curves are analyzed using the Wilson-Devinney code (W-D) for the derivation of the geometric and photometric elements of the systems. We compare the methods of photometric and spectroscopic mass ratio determination in these binaries, as a function of all typical difficulties, which arise during the analysis of such systems (light curve asymmetries, third light etc). Finally, a new spot model is suggested for the eclipsing system XY~UMa, which belongs to the RS~CVn type of active binaries. ", "introduction": "The B, V, R and I-band observations of both systems were carried out by means of CCD differential photometry. DD~Mon was observed on March 10-13, 2005 with the 1.22m Cassegrain reflector at the Kryoneri Astronomical Station of the National Observatory of Athens, Greece, while XY~UMa was observed on December 1, 4, 5 and 8, 2006 with the 0.40m Cassegrain reflector at the University of Athens Observatory, Greece. The light curves were analyzed by using the PHOEBE 0.29d software \\citep{PZ05}, which utilizes the W-D code \\citep{WD71,W79}. For the photometric light curve analysis, we performed a q-search on both systems in modes 2, 4, 5, for a rough estimation of the photometric mass ratio ($q_{ph}$). We chose the range of ${0 5)$ of $0.5$ eV would certainly have left a strong imprint on cosmology which has not been observed. In consequence, if the direct searches for a neutrino mass or the neutrinoless double beta decay indicate a present neutrino mass larger than $0.5$ eV, this could be interpreted as a strong signal in favor of a growing neutrino mass. \\\\ \\\\ \\\\" }, "0910/0910.1044_arXiv.txt": { "abstract": "We present new tests for disruption mechanisms of star clusters based on the bivariate mass-age distribution $g(M, \\tau)$. In particular, we derive formulae for $g(M,\\tau)$ for two idealized models in which the rate of disruption depends on the masses of the clusters and one in which it does not. We then compare these models with our {\\it Hubble Space Telescope\\/} observations of star clusters in the Antennae galaxies over the mass-age domain in which we can readily distinguish clusters from individual stars: $\\tau \\la 10^7 (M/10^4 M_{\\odot})^{1.3}$~yr. We find that the models with mass-dependent disruption are poor fits to the data, even with complete freedom to adjust several parameters, while the model with mass-{\\it in\\/}dependent disruption is a good fit. The successful model has the simple form $g(M,\\tau) \\propto M^{-2} \\tau^{-1}$, with power-law mass and age distributions, $dN/dM \\propto M^{-2}$ and $dN/d\\tau \\propto \\tau^{-1}$. The predicted luminosity function is also a power law, $dN/dL \\propto L^{-2}$, in good agreement with our observations of the Antennae clusters. The similarity of the mass functions of star clusters and molecular clouds indicates that the efficiency of star formation in the clouds is roughly independent of their masses. The age distribution of the massive young clusters is plausibly explained by the following combination of disruption mechanisms: (1) removal of interstellar material by stellar feedback, $\\tau \\la 10^7$~yr; (2) continued stellar mass loss, $10^7 {\\rm yr} \\la \\tau \\la 10^8 {\\rm yr}$; (3), tidal disturbances by passing molecular clouds, $\\tau \\ga 10^8$~yr. None of these processes is expected to have a strong dependence on mass, consistent with our observations of the Antennae clusters. We speculate that this simple picture also applies---at least approximately---to the clusters in many other galaxies. ", "introduction": "Star clusters are born in molecular clouds and are then dispersed into the surrounding stellar field by a variety of processes operating on different timescales, including the removal of interstellar material by stellar feedback, continued stellar mass loss, tidal disturbances by passing molecular clouds, gravitational shocks during rapid passages near the galactic bulge or through the galactic disk, orbital decay into the galactic center caused by dynamical friction, and stellar escape driven by internal two-body relaxation. Signatures of these processes are encoded in the mass function, $\\psi(M) \\propto dN/dM$, and the age distribution, $\\chi(\\tau) \\propto dN/d\\tau$, for a population of clusters in a particular galaxy. A more informative, but less familiar statistic is the bivariate distribution of masses and ages, $g(M,\\tau) \\propto \\partial^2N/\\partial{M}\\partial{\\tau}$. The key feature of $g(M,\\tau)$ is that it includes correlations between $M$ and $\\tau$, information that is absent from the univariate distributions $\\psi(M)$ and $\\chi(\\tau)$. Thus, with the help of $g(M, \\tau)$, we can answer questions such as whether low-mass clusters are disrupted faster than, or at the same rate as, high-mass clusters. In this paper, we derive formulae for $g(M, \\tau)$ for the first time for several different models for the disruption of clusters, and we then compare these formulae with our {\\it Hubble Space Telescope\\/} ({\\it HST\\/}) observations of clusters in the Antennae galaxies. In a companion paper, we present similar comparisons for the clusters in the Large and Small Magellanic Clouds (LMC and SMC; Chandar et al. 2009, hereafter CFW09). We have focused on the star clusters in the Antennae galaxies for several reasons. First, the population of clusters is large, $N \\approx 2300$ brighter than $M_V = -9$, ample for statistical studies. Second, many of the bright young clusters have masses in the range of old globular clusters, $10^4 \\la M/M_{\\odot} \\la 10^6$, suggesting that the former may simply be younger versions of the latter. Third, because the Antennae are currently in the throes of a major merger, they give us a close-up, internal view of events and processes that were more common in galaxies in the past, during their hierarchical formation. Fourth, the Antennae have been mapped at almost every wavelength available to astronomers, from radio to X-rays, providing a more detailed picture of their stellar and interstellar contents than for almost any other galaxy outside the Local Group. In our previous studies of the clusters in the Antennae, we found that the mass and age distributions could be approximated by power laws: $\\psi(M) \\propto M^{\\beta}$ with $\\beta \\approx -2$ (Zhang \\& Fall 1999, hereafter ZF99) and $\\chi(\\tau) \\propto \\tau^{\\gamma}$ with $\\gamma \\approx -1$ (Fall et al. 2005, hereafter FCW05). In these studies, we also found that the shapes of $\\psi(M)$ and $\\chi(\\tau)$ are approximately independent of the age and mass limits, respectively, of the samples used to determine them. These findings suggest that the bivariate mass-age distribution can be approximated simply by the product of the univariate distributions: \\begin{equation} g(M, \\tau) \\propto \\psi(M)\\chi(\\tau) \\propto M^{\\beta}\\tau^{\\gamma} \\end{equation} \\begin{equation} {\\rm for} \\,\\,\\,\\,\\,\\,\\,\\,\\, \\tau \\la 10^7 (M/10^4 M_{\\odot})^{1.3}\\,{\\rm yr}. \\end{equation} The second expression above indicates the approximate domain of validity of the first, the condition that objects in the sample be brighter than the most luminous individual stars ($L_V \\ga 3\\times10^5 L_{\\odot}$).\\footnote {Equation~(2) follows from the fact that the mass-to-light ratios of clusters vary with age approximately as $M/L_V \\propto \\tau^{0.8}$ for $\\tau \\ga 10^7$~yr.} In this model, clusters form with a power-law initial mass function and are then disrupted at rates that are independent of their masses. We have already explored some of the consequences of equation~(1) in two recent papers (Fall 2006; Whitmore et al. 2007, hereafter WCF07). In this paper, we make further tests of this model, and we also compare our {\\it HST\\/} observations of the Antennae clusters with two models in which clusters of different masses are disrupted at different rates. The age distribution for a population of star clusters reflects, in principle, a combination of both formation and disruption rates. We interpret the decline of $\\chi(\\tau)$ primarily as the result of disruption rather than formation for the following reasons. First, the age distribution has a sharp peak at the present time, $\\tau = 0$, and a small width, characterized by the median age $\\tau \\sim 10^7$~yr. This requires either rapid disruption at an arbitrary time (now) or rapid formation at a special time (now), the former being much more likely a priori than the latter. Of course, the Antennae galaxies are presently merging, which has almost certainly boosted the formation rates of stars and clusters, but this occurs more slowly, on the orbital timescale of the galaxies, $\\tau \\sim {\\rm few} \\times 10^8$~yr. In the simulations of merging galaxies by Mihos et al. (1993), including the Antennae, for example, the star formation rate varies by factors of only a few or less in the past $10^8$~yr, whereas the observed age distribution of the clusters drops by nearly 2 orders of magnitude in this interval of time. The second reason we interpret the decline of $\\chi(\\tau)$ in terms of disruption is that it has the same shape, including the sharp peak at $\\tau = 0$, in large regions separated by distances of order 10~kpc within the Antennae galaxies (Whitmore 2004; WCF07). There is no physical mechanism that could synchronize a burst of cluster formation this precisely at these separations; the communication time is $\\tau \\sim 10^9$~yr for a signal traveling at the typical random velocity in the interstellar medium (ISM; $v \\sim 10$~km s$^{-1}$) and $\\tau \\sim 10^8$~yr for one traveling at the highest bulk velocity ($v \\sim 100$~km s$^{-1}$). Thus, we conclude that the observed decline in the age distribution is caused mainly by disruption, at least for $\\tau \\la \\mbox{few}\\times10^8$~yr. We have speculated that equation~(1) for $g(M, \\tau)$ may also describe---approximately---the cluster populations in other galaxies (Fall 2006; WCF07). If this picture turns out to be generally valid, it will mean that star clusters form and evolve in much the same way in different galaxies, the primary variable being an overall scale factor proportional to the total number of clusters and hence to their average formation rate. We define a cluster here to be any concentrated aggregate of stars with a density much higher than that of the surrounding stellar field, whether or not it is gravitationally bound, since this is virtually impossible to determine from the available observations, especially for clusters younger than about 10 internal crossing times. Most, if not all, stars form in clusters (as just defined) and most clusters then dissolve into the stellar field. We find it intriguing that this whole complex process might be represented approximately by a simple ``recipe'' such as equation~(1). There are a few other hints that point toward this model: (1) The same mass and age distributions, $\\psi(M) \\propto M^{-2}$ and $\\chi(\\tau) \\propto \\tau^{-1}$, are found for clusters with $M \\la 10^3 M_{\\odot}$ and $\\tau \\la 3 \\times 10^8$~yr in the solar neighborhood (Lada \\& Lada 2003). (2) The luminosities of the brightest clusters in different galaxies scale approximately with the size of the population in the way expected from equation~(1) (Larsen 2002; Whitmore 2003; WCF07). (3) The mass spectrum of molecular clouds, from which the clusters form, appears to be similar in different galaxies (Blitz et al. 2007). (4) The most important early disruption processes, removal of ISM by stellar feedback and continued stellar mass loss, are independent of the properties of the host galaxy (except possibly its metallicity and stellar initial mass function (IMF)). Of course, a definitive conclusion about the generality of equation~(1) will require more detailed studies, such as that presented here for the Antennae, of the cluster populations in several other galaxies. We have recently made a start on this with the LMC and SMC (CFW09). Our main purpose in this paper is to present some of the interpretive tools needed for such studies. The plan for the remainder of this paper is the following: In \\S2, we present analytic formulae for $g(M, \\tau)$ for two models with mass-dependent disruption and the model already mentioned with mass-independent disruption. In \\S3, we describe our {\\it HST\\/} observations of the Antennae clusters and compare these with the models presented in the previous section. We also derive the luminosity function of the Antennae clusters and show how this is related to the mass function through $g(M, \\tau)$. In \\S4, we discuss the physical processes most likely responsible for the formation and disruption of clusters and how these processes affect the mass and age distributions. This interpretative section presents our current understanding of the entire life cycles of star clusters. In \\S5, we summarize our main results and their broader implications. The paper also includes two appendices. The first (A) discusses some related work on the mass and luminosity functions of the Antennae clusters by Mengel et al. (2005) and by Anders et al. (2007) respectively, while the second (B) presents some specialized formulae needed in \\S3. ", "conclusions": "In this paper, we have derived analytical formulae for the bivariate mass-age distribution $g(M,\\tau)$ for three idealized models of the formation and disruption of star clusters. We have also derived formulae for the corresponding averages of $g(M, \\tau)$ over finite intervals of $M$ and $\\tau$, denoted by $\\bar{g}(\\tau)$ and $\\bar{g}(M)$ respectively. In all three models considered here, clusters are assumed to form with a power-law initial mass function at a constant rate. In the first two models, proposed by Boutloukos \\& Lamers (2003), clusters are disrupted on a timescale that depends on their masses as $\\tau_d(M) \\propto M^k$ with $k > 0$, either suddenly (Model~1) or gradually (Model~2). In the third model, clusters are disrupted (suddenly or gradually) at a fractional rate independent of their masses, as indicated by our earlier studies of the Antennae clusters (ZF99, FCW05, WCF07). Model~3 has the remarkably simple bivariate distribution $g(M, \\tau) \\propto M^{\\beta}\\tau^{\\gamma}$. The corresponding luminosity, mass, and age distributions are all pure power laws: $\\phi(L) \\propto L^{\\alpha}$, $\\psi(M) \\propto M^{\\beta}$, $\\chi(\\tau) \\propto \\tau^{\\gamma}$. We have compared Models 1, 2, and 3 with the empirical distributions of luminosities, masses, and ages for a large sample of star clusters in the Antennae galaxies. These are based on our \\textit{UBVI}H$\\alpha$ photometry of point-like sources in images taken with the WFPC2 on {\\it HST\\/} and comparisons with stellar population models to estimate $M$ and $\\tau$ for each source (after correcting for interstellar reddening). To distinguish clusters of stars from individual stars, we have restricted our sample for analysis to objects brighter than $M_V = -9$, corresponding to $L > 3 \\times 10^5 L_{\\odot}$. This limit defines the domain of validity of the models in the $L$-$\\tau$ and $M$--$\\tau$ planes, the regions above the solid curves in Figures~1 and 2 respectively, corresponding approximately to $\\tau \\la 10^7 (M/10^4 M_{\\odot})^{1.3}$~yr. Thus, all the results presented in this paper pertain to relatively massive and relatively young clusters (except in \\S4.3). We find that Model~3, with mass-{\\it in\\/}dependent disruption, provides a good match to the observed mass-age distribution of the Antennae clusters. The best-fitting exponents in this model are $\\beta \\approx -2$ and $\\gamma \\approx -1$. Models~1 and 2, with mass-dependent disruption fare much worse. Even with complete freedom to adjust several parameters, these models never come close to matching the data. While our detailed comparisons are based on models in which the formation rates of clusters are constant, our main conclusion, that the disruption rates do not depend on mass, is valid even if the formation rates are variable. As a check on these results, we have also rederived the luminosity function of the Antennae clusters. After corrections for interstellar extinction, we find a good fit to a pure power law, $\\phi(L) \\propto L^{\\alpha}$, with $\\alpha \\approx -2$. The fact that the mass and luminosity functions are nearly identical power laws ($\\alpha \\approx \\beta$) is a consequence of---and additional support for---the weak or nonexistent correlations between masses and ages, i.e., the decomposition $g(M,\\tau) \\propto \\psi(M)\\chi(\\tau)$. We have also investigated recent claims that there are bends or other features in the mass and luminosity functions of the Antennae clusters (Fritze-von Alvensleben 1999; Mengel et al. 2005; Anders et al. 2007). We show that these features are artifacts of the ways the samples are defined and do {\\it not\\/} reflect physical processes involved in the formation and disruption of the clusters. In an effort to understand the physical basis for our simple---but doubtless approximate---model for $g(M,\\tau)$, we first note that the resulting mass function, $\\psi(M) \\propto M^{-2}$, resembles that of molecular clouds in nearby galaxies. This indicates that the efficiency of star formation in the clouds is roughly independent of their masses. We also consider a variety of processes that could disrupt the protoclusters and clusters: removal of ISM by stellar feedback, continued stellar mass loss, tidal disturbances by passing molecular clouds, gravitational shocks during rapid passages near the galactic bulge and through the galactic disk, orbital decay into the galactic center caused by dynamical friction, and stellar escape driven by internal two-body relaxation. We suggest that the massive young clusters in the Antennae galaxies are disrupted in the following approximate sequence: (1) ISM removal, $\\tau \\la 10^7$~yr, (2) stellar mass loss, $10^7 {\\rm yr} \\la \\tau \\la 10^8 {\\rm yr}$, (3) tidal disturbances, $\\tau \\ga 10^8$~yr. We have argued on theoretical grounds that these processes would operate at rates roughly independent of the masses of the clusters, consistent with our empirically-based decomposition $g(M, \\tau) \\propto \\psi(M)\\chi(\\tau)$ (which, however, is only well established for $\\tau \\la 10^8$~yr). In combination, these processes plausibly account for the observed decline of the age distribution, $\\chi(\\tau) \\propto \\tau^{-1}$. In the longer term, after 12 Gyr or so, the escape of stars driven by two-body relaxation will preferentially disrupt low-mass clusters and imprint a peak or turnover in the evolved mass function at $M_p \\approx (1-2) \\times 10^5 M_{\\odot}$, similar to that for old globular clusters in the Milky Way and other galaxies. How general is this picture? In the Introduction, we mentioned some circumstantial evidence in favor of its wide applicability: the statistics of embedded clusters in the solar neighborhood (Lada \\& Lada 2003), the luminosities of the brightest clusters in different galaxies (Larsen 2002; Whitmore 2003; WCF07), the similarity of the mass spectra of molecular clouds in different galaxies (Blitz et al. 2007), and the fact that the early disruption of protoclusters and clusters ($\\tau \\la 10^8$~yr) is driven mainly by internal processes that depend weakly, if at all, on the properties of their host galaxies (\\S4 here). Furthermore, from a detailed study of the mass-age distributions of clusters in the LMC and SMC, we find results nearly identical to those presented here for the clusters in the Antennae (CFW09). This is a strong test because the Antennae and the Magellanic Clouds represent very different environments for star and cluster formation: two large interacting galaxies on one hand, two small, relatively quiescent galaxies on the other hand. Nevertheless, further tests of this picture in other galaxies would be beneficial. If this picture turns out to be generally valid, it will mean that the main difference between populations of young clusters is simply in the normalization of $g(M, \\tau)$ and hence in the overall formation rate. From this perspective, the objects traditionally designated open, populous, globular, or super clusters would belong to a continuum, with mass and age as the most fundamental variables, and would not require different formation and/or disruption processes to account for their observed properties. The picture presented here has the virtues of simplicity and possible universality. It captures the salient properties of the observed mass-age distribution, at least for the populations of clusters we have studied so far. Nevertheless, we do not expect our model $g(M, \\tau)$ to be strictly universal. It may break down outside its domain of validity in most galaxies or inside this domain in some extreme environments, such as those dominated by population III stars. There may also be minor deviations from the model in some normal galaxies, even within the domain of validity, caused for example by variations in the formation rates and/or differences in the disruption rates by passing molecular clouds. However, these are only likely to affect $g(M, \\tau)$ at intermediate ages ($10^8 {\\rm yr} \\la \\tau \\la 10^9{\\rm yr}$), which will be difficult to detect observationally. We see an analogy between, on one hand, the potentially universal stellar IMF, represented by a simple, approximate formula (first by Salpeter (1955) and then revised in subsequent studies based on more modern data) and, on the other hand, the potentially universal mass-age distribution $g(M, \\tau)$ for star clusters, represented here by a simple, approximate formula. Both the stellar IMF and the cluster $g(M, \\tau)$ are thought to be the outcome of several complex physical processes, in ways not yet fully understood, and neither is expected to hold exactly in all situations. Even so, these functions provide compact and useful summaries of the observations and are thus natural focal points for future studies of the formation and early evolution of stars and clusters." }, "0910/0910.3277_arXiv.txt": { "abstract": "Imaging atmospheric Cherenkov telescopes (IACTs) used for ground-based gamma-ray astronomy at TeV energies use reflectors with areas on the order of 100m$^2$ as their primary optic. These tessellated reflectors comprise hundreds of mirror facets mounted on a space frame to achieve this large area at a reasonable cost. To achieve a reflecting surface of sufficient quality one must precisely orient each facet using a procedure known as alignment. We describe here an alignment system which uses a digital (CCD) camera placed at the focus of the optical system, facing the reflector. The camera acquires a series of images of the reflector while the telescope scans a grid of points centred on the direction of a bright star. Correctly aligned facets are brightest when the telescope is pointed directly at the star, while mis-aligned facets are brightest when the angle between the star and the telescope pointing direction is twice the misalignment angle of the facet. Data from this scan can be used to calculate the adjustments required to align each facet. We have constructed such a system and have tested it on three of the VERITAS IACTs. Using this system the optical point spread functions of the telescopes have been narrowed by more than 30\\%. We present here a description of the system and results from initial use. ", "introduction": "The current generation of imaging atmospheric Cherenkov telescopes operating around the world~\\cite{Holder08,Hinton04,Baixeras04} has ushered in a new era in TeV gamma-ray astronomy. The number of detected TeV gamma-ray sources has grown from below ten in 2000 to more than seventy today~\\cite{Aharonian08} largely because of the increased sensitivity of the instrumentation. This increase results from the use of the following: \\begin{itemize} \\item{larger reflectors} \\item{cameras with larger fields of view and higher resolution} \\item{multiple telescopes making stereoscopic observations} \\item{flash-ADC-based data acquisition systems} \\end{itemize} For the benefits afforded by large reflectors and high resolution cameras to be fully realised, the optical quality of the telescopes must be maintained at a high level. Since the reflectors of these telescopes comprise several hundred mirror facets, their alignment presents a significant technical and logistical challenge. The VERITAS array, located in southern Arizona, USA, employs four twelve-metre-diameter f-1.0 reflectors of the Davies-Cotton type~\\cite{DaviesCotton}. Each reflector consists of a tubular steel optical support structure (OSS) on which 345 identical hexagonal mirror facets are mounted. The facet mounts allow precision adjustments to bring the focus of each to the same point on the primary focal plane of the telescope. Since the mirror facets are exposed to the dust of the Arizona Sonoran desert their reflectivity degrades over time so they are therefore re-coated on a regular basis~\\cite{Roache08}. This process maintains the reflectivity of the facets but their removal and re-installation compromises the optical quality of the reflector as a whole so the alignment of the facets must be repeated on a regular basis. We present here an alignment system, based on the technique originally suggested by Arqueros et al.~\\cite{Arqueros05}, which can be used for aligning the VERITAS telescopes. It achieves the quality desired in a reasonable length of time at modest cost. Importantly, the optimal alignment is achieved for typical observation elevations. ", "conclusions": "An alignment system based on the technique suggested by \\cite{Arqueros05} has been developed and used to improve the optics of three VERITAS telescopes. This has led to a reduction in the size of the PSF by more than 30\\%. Moreover this system is less labour-intensive than that which was previously used. It has the advantage that the telescope reflectors are directly optimised for use at typical observing elevations. Further investigations are planned." }, "0910/0910.2792_arXiv.txt": { "abstract": "We present the detailed optical to far-infrared observations of SST J1604+4304, an ULIRG at $z = 1.135$. Analyzing the stellar absorption lines, namely, the CaII H \\& K and Balmer H lines in the optical spectrum, we derive the upper limits of an age for the stellar population. Given this constraint, the minimum $\\chi^2$ method is used to fit the stellar population models to the observed SED from 0.44 to 5.8\\micron. We find the following properties. The stellar population has an age 40 - 200 Myr with a metallicity 2.5 $Z_{\\sun}$. The starlight is reddened by $E(B-V) = 0.8$. The reddening is caused by the foreground dust screen, indicating that dust is depleted in the starburst site and the starburst site is surrounded by a dust shell. The infrared (8-1000\\micron) luminosity is $L_{ir} = 1.78 \\pm 0.63 \\times 10^{12} L_{\\sun}$. This is two times greater than that expected from the observed starlight, suggesting either that 1/2 of the starburst site is completely obscured at UV-optical wavelengths, or that 1/2 of $L_{ir}$ comes from AGN emission. The inferred dust mass is $2.0 \\pm 1.0 \\times 10^8 M_{\\sun}$. This is sufficient to form a shell surrounding the galaxy with an optical depth $E(B-V) = 0.8$. From our best stellar population model - an instantaneous starburst with an age 40 Myr, we infer the rate of 19 supernovae(SNe) per year. Simply analytical models imply that 2.5 $Z_{\\sun}$ in stars was reached when the gas mass reduced to 30\\% of the galaxy mass. The gas metallcity is $4.8 Z_{\\sun}$ at this point. The gas-to-dust mass ratio is then $120 \\pm 73$. The inferred dust production rate is $0.24 \\pm 0.12 M_{\\sun}$ per SN. If 1/2 of $L_{ir}$ comes from AGN emission, the rate is $0.48 \\pm 0.24 M_{\\sun}$ per SN. We discuss the evolutionary link of SST J1604+4304 to other galaxy populations in terms of the stellar masses and the galactic winds, including optically selected low-luminosity Lyman $\\alpha$-emitters and submillimeter selected high-luminosity galaxies. ", "introduction": "$IRAS$(Infrared Astronomical Satellite) observations discovered numerous infrared galaxies. At bolometric luminosities $> 10^{11} L_{\\sun}$, infrared galaxies are the dominant population in the local Universe, being more numerous than optically selected starbursts, AGNs, and quasars \\citep{sanders96}. By luminosity, infrared galaxies are classified into luminous infrared galaxies(LIRGs) with $L_{ir} > 10^{11} L_{\\sun}$, ultraluminous infrared galaxies(ULIRGs) with $L_{ir}> 10^{12} L_{\\sun}$, and hyperluminous infrared galaxies(HyLIRGs) with $L_{ir}> 10^{13} L_{\\sun}$. The ratio of infrared to visible luminosity, $L_{ir}/L_B$ increases with $L_{ir}$ \\citep{soifer89}. Although there is evidence that an optically buried AGN may exist in LIRGs, there is similarly strong evidence that enhanced star formation is ongoing (\\citealt{armus95}; \\citealt{sanders96}). \\citet{heckman90} found that their optical spectroscopic data support the superwind model in which the kinetic energy provided by supernovae (SNe) and winds from massive stars in starburst site drives a large-scale outflow. The {\\it Infrared Space Observatory (ISO)} source counts in the mid- and far-infrared extended our knowledge of infrared galaxies up to $z \\sim 1$ (\\citealt{taniguchi97};\\citealt{oliver97};\\citealt{kawara98}; \\citealt{puget99}; \\citealt{elbaz99};\\citealt{serjeant00}; \\citealt{sato03}; \\citealt{kawara04}),and revealed the number of ULIRGs rapidly grew with increasing redshift(\\citealt{genzel00}; \\citealt{chary01}; \\citealt{heraudeau04}; \\citealt{oyabu05}). The next large advance was made by the {\\it Spitzer Space Telescope} with the sensitivity to probe infrared galaxies at $z = 2- 4$ (see review in \\citealt{soifer08} reference therein). \\citet{reddy06}, using deep {\\it Spitzer} 24\\micron\\ observations, examined star formation and extinction in optically selected galaxies, near-infrared-selected galaxies, and submillimeter-selected galaxies (SMGs) at $z \\sim 2$. {\\it Spitzer} provided a powerful probe to measure redshifts of SMGs, the most luminous infrared galaxy population as HyLIRGs \\citep{pope08a}. Based on {\\it Spitzer} mid-infrared data, the diagnostic diagram to separate starburst- and AGN-dominated infrared galaxies has been developed(e.g., \\citealt{stern05}; \\citealt{pope08b}; \\citealt{polletta08}), and it is suggested that $L_{ir}$ in SMGs as well as ULIRGs at $z \\sim 2$ is dominated by star formation\\citep{pope08a}, while many SMGs contain an AGN as evidenced by their X-ray properties. \\citet{dey08} uncovered a new population of dust-obscured galaxies (DOGs). These galaxies have $L_{ir}$ comparable to ULIRGs. However, their $L_{ir}/L_B$ is greater than local ULIRGs. DOGs and SMGs as well as high-redshift ($z \\sim 1$) ULIRGs are generally considered to be young and massive galaxies, although the nature of the stellar populations is far from being understood. To understand the evolutionary link of these infrared galaxies to optically selected and near-infrared-selected galaxies, it is crucial to observe stellar absorption lines which indicate the age of the stellar populations. During the course of searching for distant objects in a cluster field, SST J1604+4304 drew our attention with its red optical-3.6\\micron\\ color. Follow-up study revealed this object is an ULIRG at $z \\sim 1.135$ and its energy source is young stars with an age $< 200$ Myr. We present the data consisting of optical spectroscopy and photometry from the optical to far-infrared in section 2. We derive an age and metallicity of the stellar population, and show a foreground dust screen is plausible in Section 3. In Section 4, mid- and far-infrared emission by dust and energetics are discussed. In Section 5, we discuss the inferred size of a dust shell surrounding the galaxy, the SN rate and metallicity expected from the broad-band SED analysis, and the dust production rate per SN. The evolutionary link of infrared galaxies to optically selected galaxies is discussed based on the galactic wind models of elliptical galaxies. We adopt $H_0$ = 70 km s$^{-1}$ Mpc$^{-1}$, $\\Omega_m = 0.3$, and $\\Omega_{\\Lambda} = 1 - \\Omega_m$ throughout this paper. In this cosmology, the luminosity distance to the object is 7520 Mpc. The flux density $F_{\\nu}$ is used in units of $\\mu$Jy, which is converted to AB magnitudes through the relation $m(AB) = -2.5 log(F_{\\nu})$ + 23.9. \\section[]{Observations and data analysis} This work is based on the archival data taken by the {\\it Spitzer Space Telescope}, the {\\it Hubble Space Telescope (HST)}, and the $Subaru$ Telescope and new photometric observations performed on the {\\it United Kingdom Infrared Telescope(UKIRT)} and the $MAGNUM$\\footnote{This telescope is dedicated to the Multicolor Active Galactic Nuclei Monitoring (MAGNUM) project and is installed at the Haleakala Observatories in Hawaii \\citep{yoshii02}.} telescope. In addition, the {\\it Gemini Telescope North} was used to do optical spectroscopy. \\begin{figure} \\includegraphics[width=84mm]{k3_chart.ps} \\caption{Images centered at SST J1604+4304, where the size is 70\\arcsec $\\times$ 70\\arcsec for $I_{814}$, 3.6\\micron, 8.0\\micron, and 24\\micron, and 240\\arcsec $\\times$ 240\\arcsec for 70\\micron, and 160\\micron. For all plots, north is at the top and east is to the left. We present 32\\arcsec radius circles which are the size of the aperture used for MIPS 160\\micron\\ photometry. Two bars in the $I_{814}$ image are used to indicate SST J1604+4304. } \\label{f_chart} \\end{figure} \\begin{figure} \\includegraphics[width=84mm]{k3_closeup.ps} \\caption{10\\arcsec $\\times$ 10\\arcsec images centered at SST J1604+4304. For all plots, north is at the top and east is to the left. We present 3\\farcs7 radius circles which are the size of the aperture used for IRAC 3.6-8.0\\micron\\ photometry. } \\label{f_closeup} \\end{figure} \\subsection{Target} SST J1604+4304 has been found at \\\\ \\begin{flushleft} R.A. 16$^h$ 04$^m$ 25\\fs538\\\\ decl. +43\\degr 04\\arcmin 26\\farcs55,\\\\ \\end{flushleft} \\noindent in the J2000 system in a field observed with the $Spitzer$ IRAC instrument. The optical counterpart in the $HST$ data is located within the 0\\farcs4 offset accuracy requirements of $Spitzer$ \\citep{werner04}. Images of SST J1604+4304 are given in Figures \\ref{f_chart} and \\ref{f_closeup}. The broad-band fluxes are given in Table \\ref{t_fluxes}. SST J1604+4304 is only 32\\arcsec away from the center of the massive galaxy cluster CL 1604+4304 at $z = 0.90$ \\citep{gal04}. Based on a weak-lensing analysis, \\citet{margoniner05} find the mass contained within projected radius $R$ is $(3.69 \\pm 1.47)[R/(500 kpc)] \\times 10^{14} M_{\\sun}$, corresponding to an inferred velocity dispersion 1004 $\\pm$ 199 km s$^{-1}$. Thus, using a singular isothermal sphere, we obtain the magnification $m = 1.17 \\pm 0.09$ for SST J1604+4304 \\citep{schneider92}. \\subsection{$Spitzer$ data} All IRAC and MIPS data sets were retrieved from the $Spitzer$ public archive. The IRAC data consist of four broadband images at 3.6, 4.5, 5.8, and 8.0 \\micron. The field-of-view is nearly 5\\farcm2 $\\times$ 5\\farcm2 with a linear scale of approximately 1\\farcs22 pixel$^{-1}$. The reader is referred to \\citet{fazio04} for a complete report on the in-flight performance of IRAC. In the pipeline processing, ten BCD FITS images were co-added into a single mosaic image per band with a technique similar to 'drizzle'. The total effective exposure time is 1936 sec pixel$^{-1}$ per band. Photometry was performed using a 3 pixel (3\\farcs7) radius aperture. As can be seen in the 3.6\\micron\\ image in Figure \\ref{f_chart}, there are two sources which can contribute to the flux within the photometric aperture; a faint source is at 6\\farcs5 to the southwest and a bright source 10\\arcsec to the west. Using the point-spread function which is available on the IRAC website, their contributions were estimated to be 3.1 - 4.3 \\% of the flux of SST J1604+4304 depending on the photometric band. The total fluxes after applying the aperture correction are given in Table \\ref{t_fluxes}. The quoted errors include 5\\% absolute calibration uncertainty that SSC\\footnote{See Infrared Array Camera Data Handbook} recommends for general observers. Note that the color correction is very small and ignored. The MIPS data were taken at 24, 70, and 160 \\micron\\ in 2004 March. In the pipeline processing, the mapped BCD data were rebinned and coadded to a single image with a linear pixel scale of 2\\farcs45, 4\\farcs0, and 4\\farcs0 pixel$^{-1}$ for 24, 70, and 140 \\micron, respectively. The total integration times are 93, 53, and 84 sec pixel$^{-1}$. In the following analysis, we use the following post-BCD products, namely, mosaic images for 24\\micron\\ and filtered mosaic images for 70 and 160\\micron. 24\\micron\\ photometry was performed using a 7\\arcsec radius aperture. The contributions, to the flux within the photometric aperture, from the two sources (see the 24\\micron\\ image in Figure \\ref{f_chart}), located approximately at 15\\arcsec to the west-northwest and 13\\arcsec to the southwest, were estimated using the point-spread function on the MIPS website. Their contributions are 9\\% of the flux within the aperture. Then, the color correction for 500K blackbody and the aperture correction were applied. MIPS 70 and 160 \\micron\\ photometry was performed using 16\\arcsec and 32\\arcsec radius apertures, respectively. As seen in Figure \\ref{f_chart}, SST J1604+4304 is faint in the far-infrared. We measured sky fluxes with the same aperture at 10 or 11 positions along the annulus with radii of 36\\arcsec - 72\\arcsec at 70\\micron\\ and 96 - 150\\arcsec at 160\\micron. The flux averaged over these sky fluxes was subtracted from the flux measured within the object aperture centered at SST J1604+4304. The aperture correction and the color correction for 30K were applied. The detections are 2.5$\\sigma$ and 5.0$\\sigma$ at 70 and 160\\micron, respectively. As seen in the 24\\micron\\ image in Figure \\ref{f_chart}, there are several 24\\micron\\ sources within the 160\\micron\\ photometry aperture. Judging from their [3.6\\micron]-[24\\micron] color and 70 and 160 \\micron\\ brightness at their respective positions, the source at 13\\farcs5 to the west-northwest, the only source as red as SST J1604+4304, may contribute 1/3 of the flux within the aperture at most. The 160 \\micron\\ flux is 38.4 $\\pm$ 7.65 mJy if the contributions from this source can be ignored, while the flux is 25.6 $\\pm$ 7.65 mJy if this source contributes 1/3 of the flux within the aperture. The real flux of SST J1604+4304 should be between them, and thus the flux is 32 $\\pm$ 14 mJy. The aperture correction and the color correction for 30K blackbody were applied. Note that the 24\\micron\\ error in Table \\ref{t_fluxes} includes 4\\% absolute calibration uncertainty\\footnote{Multiband Imaging Photometer for Spitzer (MIPS) Data Handbook}. \\subsection{Optical imaging data} The $V_{606}$, $I_{814}$, and $Z_{850}$ image data were retrieved from the $HST$ public archive, where $V_{606}$, $I_{814}$, and $Z_{850}$ denote the $F606W$, $F814W$, and $F850LP$, respectively. The Advanced Camera for Surveys (ACS) Wide Field Camera (WFC) was used, which covers a field of 3\\farcm4 $\\times$ 3\\farcm4 with a scale of 0\\farcs05 pixel$^{-1}$. Four image data were combined into a single co-added image per band. The total exposure times are 4840 sec pixel$^{-1}$ for $V_{606}$ and $I_{814}$ and 6000 sec pixel$^{-1}$ for $Z_{850}$. The $I_{814}$ image in Figure \\ref{f_closeup} shows that SST J1604+4304 has irregular morphology extending 1\\farcs6 in the NW direction with a 0\\farcs6 width. To determine the total flux of this object, $I_{814}$ photometry was performed using 0\\farcs8, 1\\farcs0, 1\\farcs2, and 2\\farcs0 radius apertures. The $I_{814}$ flux grows from 0\\farcs8 to 1\\farcs2, while it barely increase from 1\\farcs2 to 2\\farcs0. Note that contributions from the two nearest sources were subtracted from the flux measured with the 2\\farcs0 radius aperture. Thus, the flux measured with the 1\\farcs2 radius aperture represents the total flux of SST J1604+4304. The ratio of the flux with a 0\\farcs8 aperture to a 1\\farcs2 aperture is 1.24 $\\pm$ 0.05. $V_{606}$ and $Z_{850}$ photometry was performed using a 0\\farcs8 aperture. The resultant fluxes were converted to the total fluxes by multiplying 1.24 $\\pm$ 0.05, because we did not measure significant changes in optical colors across the target. The total fluxes are given in Table \\ref{t_fluxes}. The errors in Table \\ref{t_fluxes} include the error of aperture conversion from 0\\farcs8 - 1\\farcs2 and the correction for correlated noise in drizzling. $Subaru$ $B, V, R_c, I_c, z$ data were retrieved from the SMOKA science achieve. The data were taken with Suprime-Cam (SUP) imager which has a field-of-view of 34\\arcmin $\\times$ 27\\arcmin with a scale of 0\\farcs2 pixel$^{-1}$. The total exposure time is 2880, 2520, 3600, 1680, 3300 sec pixel$^{-1}$ after co-adding 4 to 11 image frames, using the standard SUP script \\citep{yagi02}. Flux scaling was performed observing 11 standard stars given by \\citet{majewski94}. $I_c$ photometry were performed using three apertures with a radius of 1\\farcs0, 1\\farcs2, and 2\\farcs0, and it was confirmed the $Subaru$ $I_c$ brightness distribution is identical to that measured at the $HST$ $I_{814}$ photometry band. Thus, we use the 1\\farcs2 radius aperture centered on SST J1604+4304 to obtain the fluxes given in Table \\ref{t_fluxes}. The errors in Table \\ref{t_fluxes} include 5\\% photometric uncertainty. \\subsection{H/K imaging observations} The $H$-band image was taken with the UIST (1-5 micron imager spectrometer) on $UKIRT$ in February 2006. UIST was used in the imaging mode with a field-of-view of 2\\arcmin $\\times$ 2\\arcmin and a scale of 0\\farcs12 pixel$^{-1}$. The total exposure time is 1800 sec pixel$^{-1}$ after co-adding individual image data with a integration time of 60 sec. The sky condition was photometric with 0\\farcs6 seeing. The K-band image was taken with the multicolor imaging photometer (MIP) on the 2 m $MAGNUM$ telescope (\\citealt{kobayashi98};\\citealt{minezaki04}) in August 2006. MIP has a field-of-view of 1\\farcm5 $\\times$ 1\\farcm5 with a linear scale of 0\\farcs346 pixel$^{-1}$. The total integration time is 2340 sec pixel$^{-1}$ after co-adding individual image data. The weather condition was photometric with 1\\arcsec seeing. Flux calibration was performed using 2MASS 16042631+4303413 (H=14.36, K= 14.39) observed simultaneously with SST J1604+4304. Photometric uncertainty is 5\\% and 9\\% for H and K, respectively. \\begin{table} \\centering \\caption{Multiwavelength photometry of SST J1604+4304.} \\begin{tabular}{crrrr} \\hline Filter$^a$ & Total flux & $R_{ph}$$^b$ & Instrument & UT \\\\ band & $\\mu$Jy & arcsec & & yy/mm \\\\ \\hline $V_{606}$ & 0.224 $\\pm$ 0.062 & 1.2 & $HST$ ACS & 03/08 \\\\ $I_{814}$ & $1.05 \\pm 0.08$ & 1.2 & $HST$ ACS & 03/08 \\\\ $Z_{850}$ & $1.53 \\pm 0.13$ & 1.2 & $HST$ ACS & 07/02 \\\\ 3.6\\micron & $60.35 \\pm 3.16$ & 3.7 & $Spitzer$ IRAC & 04/03 \\\\ 4.5\\micron & $53.76 \\pm 2.79$ & 3.7 & $Spitzer$ IRAC & 04/03 \\\\ 5.8\\micron & $55.11 \\pm 4.07$ & 3.7 & $Spitzer$ IRAC & 04/03 \\\\ 8.0\\micron & $65.68 \\pm 5.36$ & 3.7 & $Spitzer$ IRAC & 04/03 \\\\ 24\\micron & $319 \\pm 18.4$ & 7.0 & $Spitzer$ MIPS & 04/03 \\\\ 70\\micron & $2150 \\pm 870$ & 16 & $Spitzer$ MIPS & 04/03 \\\\ 160\\micron & $32000 \\pm 14000$ & 32 & $Spitzer$ MIPS & 04/03 \\\\ \\hline $B$ & $0.0803 \\pm 0.013$ & 1.2 & $Subaru$ SUP & 00/06 \\\\ $V$ & $0.111 \\pm 0.015$ & 1.2 & $Subaru$ SUP & 05/05 \\\\ $R_c$ & $0.218 \\pm 0.018$ & 1.2 & $Subaru$ SUP & 01/04,05 \\\\ $I_c$ & $0.689 \\pm 0.042$ & 1.2 & $Subaru$ SUP & 01/04 \\\\ $z$ & $2.29 \\pm 0.21$ & 1.2 & $Subaru$ SUP & 01/06 \\\\ $H$ & $5.40 \\pm 1.19$ & 1.2 & $UKIRT$ UIST & 06/02 \\\\ $K$ & $20.06 \\pm 3.35$ & 1.2 & $MAGNUM$ MIP & 06/08 \\\\ \\hline \\hline 20cm & $< 0.66$ mJy$^c$ & 2.7 & $VLA$$^e$ & 1995 \\\\ 0.2-2 keV & $<2.1$ $^d$ & 15 & $XMM$$^f$ & 02/02 \\\\ 2-10 keV & $<15$ $^d$ & 15 & $XMM$$^f$ & 02/02 \\\\ \\hline \\end{tabular} \\medskip \\\\ \\begin{flushleft} $^a$ $V_{606}, I_{814}, \\& Z_{850}$ denote F606W,F814W,\\& F850LP, respectively.\\\\ $^b$ The radius of the photometric apertures which were used to obtain the total fluxes.\\\\ $^c$ To 3 $\\sigma$.\\\\ $^d$ To 3 $\\sigma$ in units of $10^{-15}$ erg cm$^{-2}$ s$^{-1}$.\\\\ $^e$ From the FIRST survey \\citep{white97}.\\\\ $^f$ Observed by \\citet{lubin04}. The FLIX online server was used to derive the upper limit. \\end{flushleft} \\label{t_fluxes} \\end{table} \\subsection{Optical spectroscopy} Optical spectroscopy was carried out using the Gemini Multi-Object Spectrograph (GMOS) on the Gemini-North telescope in 2007 August. The R150\\_G5306 grating was used along with the OG151\\_G036 filter, covering 5,000 - 10,000 \\AA\\ simultaneously with 6.96 \\AA/pixel after binned by 4 in the dispersion direction. The different wavelength centers (8,200 and 8,250 \\AA) were set to fill the CCD gaps. A 1\\farcs0 width slit was set so that the dispersion direction is perpendicular to the long axis of the target. Seven 1800 sec exposures were taken in the Nod \\& Shuffle mode. Data were reduced in a standard manner using the IRAF GEMINI.GMOS package. The sky subtraction was made with the GNSSKYSUB routine. Wavelengths were scaled with CuAr spectra and the relative photometric calibration was performed using the standard star (HZ44). The results are shown in the bottom panel of Figure \\ref{f_sed}. The spectrum which is median-smoothed with a width of 40 pixles (i.e., $\\sim$ 280 \\AA), is compared with the photometry from the images in the top panel of Figure \\ref{f_sed}. The two measurements agree with each other. \\begin{figure} \\includegraphics[width=84mm]{k3_sed.ps} \\caption{The top panel shows UV to far-infrared SED for SST J1604+4304. The broad-band data are shown along with the median-smoothed optical spectrum ({\\it gray}) taken with Gemini-North GMOS. The bottom panel shows the optical spectrum ({\\it top}) together with the synthetic spectrum ({\\it bottom}) at an age of 400 Myr, reddened by E(B-V) = 0.6} \\label{f_sed} \\end{figure} ", "conclusions": "For simplicity, we temporarily assume that SST J1604+4304 is purely powered by star formation and at the end we will discuss the impact if SST J1604+4304 is AGN-dominated. \\subsection{Stellar mass} The stellar mass $M_*$ can be estimated as $L_{ir}/(L/M_*)$, where $L/M_*$ is the luminosity to mass ratio for the stellar population models. Thus, we have $M_* = 6 - 13 \\times 10^{10} M_{\\sun}$ at $ t = 0$ as listed in Table \\ref{t_properties}. If this stellar mass was constantly built up during 40 - 200 Myr, the star formation rate is $SFR$ = 600 - 1650 M$_{\\sun}$ yr$^{-1}$. Using the calibrations of SFR in terms of the [OII] luminosity (\\citealt{kennicutt98}), after dereddening the [OII] flux, implies 13 - 500 $M_{\\sun}$ yr$^{-1}$. Considering that we are only looking at 1/2 of the starburst activity in the the optical, we obtain SFR = 26 - 1000 $M_{\\sun}$ yr$^{-1}$, in good agreement with the previous estimate. \\subsection{Comparison with other infrared galaxy populations} SST J1604+4304 is a dusty, super metal-rich, young galaxy at $z = 1.135$ with $E_{B-V}$ = 0.8, metallicity 2.5 $Z_{\\sun}$, and an age 40 - 200 Myr. The age is comparable to young, optical-selected populations like Lyman-$\\alpha$ emitters (LAEs) and Lyman break galaxies (LBGs). Studying the $z \\sim 3$ LAE sample, \\citet{gawiser06} inferred ages of the order of 100 Myr with stellar masses $5 \\times 10^8 M_{\\sun}$ and no dust extinction. \\citet{pirzkal07} showed $4 < z < 5.7$ LAEs had very young ages a few Mys with stellar masses $10^6 - 10^8 M_{\\sun}$. \\citet{sawicki98}, studying $z > 2$ LBGs, found that their LBGs were dominated by young stellar populations with ages $<$ 200 Myr with stellar masses several $\\times 10^9 M_{\\sun}$ and moderate dust extinction typically $E(B-V) \\sim 0.3$. The $z = 2 - 3.5$ sample of LBGs by \\citet{papovich01} has ages of 30 Myr to 1 Gyr with stellar masses $10^9 - 10^{11} M_{\\sun}$ with $E(B-V)$ = 0.0 - 0.4. SST J1604+4304, having a stellar mass 6 - 13 $\\times 10^{10} M_{\\sun}$, is more massive and more dusty than these optically selected galaxies. SST J1604+4304 having $f_{\\nu}(24\\mu m)/f_{\\nu}(R) > 1000$ meets the criteria for 24\\micron-selected high-redshift ($z \\sim 2$) dust-obscured galaxies (DOGs). DOGs and submillimeter-selected galaxies (SMGs) have extremely large $L_{ir}/L_B$ which cannot be accounted for by redshifting local ULIRGs (\\citealt{dey08};\\citealt{pope08b}). \\citet{dey08} found nearly all of DOGs are ULIRGs with $L_{IR} > 10^{12} L_{\\sun}$ and more than half of these have $L_{IR} > 10^{13} L_{\\sun}$. The infrared luminosity of radio-identified SMGs ranges $L_{IR} = 2 \\times 10^{11} - 10^{14} L_{\\sun}$ \\citep{chapman05}. A SFR 1000 $M_{\\sun} yr^{-1}$, which corresponds to $L_{IR} = 5 \\times 10^{12} L_{\\sun}$ \\citep{kennicutt98}, builds up a stellar mass of $10^{11} M_{\\sun}$ in 100 Myr. Thus, it is considered that DOGs and SMGs are progenitors of present-day massive galaxies. If ages of stellar populations in DOGs and SMGs are derived, as we did for SST J1604+4304 by measuring the stellar absorption lines, the evolutionary link of DOGs and SMGs to normal ULIRGs and optically selected galaxies will be better understood. \\subsection{SNe for dust and metal production} In SST J1604+4304 with an age 40 - 200 Myr, most of the dust would be formed by type II supernovae (SNe II), because low-mass stars had not evolved to form dust in the expanding envelope. This may be the reason why our best model results in poor goodness of fit $\\chi_{\\nu}^2(min) = 5.2$. The Calzetti's extinction law may not be applicable to such young galaxies, because dust grains in local starbursts would be dominated by those formed in the outflows of low-mass stars and not by SNe II dust. In this regard, it is of great interest to use the SED of SST J1604+4304 to test extinction laws such as those calculated by Hirashita et al. (2005, 2008) that are predicted for dust formed in SNe II ejecta \\citep{nozawa03}. Our best model is the instantaneous-burst model with an age 40 Myr with a stellar mass $6.4 \\pm 2.3 \\times 10^{10} M_{\\sun}$ at $t = 0$. The metallicity is 2.5 $Z_{\\sun}$ and the dust mass is $M_{dust} = 2.0 \\pm 1.0 \\times 10^8 M_{\\sun}$. Until 40 Myrs after the onset of the starburst, stars with a mass $> 8 M_{\\sun}$ died with a SN explosion. The number of stars evolved into SNe in 40 Myr is $7.6 \\times 10^8$ or 19 SNe per year. Adopting the nucleosynthesis models by \\citet{nomoto06}, the total yield is 0.09 - 0.11 and the metal yield is 0.016 - 0.020 for the Chabrier IMF.\\footnote{The original yields are given for the Salpeter IMF. These are converted to the Chabrier IMF by simply multiplying 1.64 to adjust the difference in the mass fraction of stars from 8 - 100 $M_{\\sun}$ to those from 0.1 -100 $M_{\\sun}$.} The gas and stetllar metallicities can be estimated applying simply analytical models based on the instantaneous recycling approximation. The gas metallicity is expressed as $Z_g(t)$ = $y(t)ln(1/f(t))$ where $f(t)$ is the ratio of the gas mass to the galaxy mass, and $y(t)$ is the ratio of the rate at which the metal is produced by the events of nucleosysnthesis and ejected into the interstellar gas to the rate at which hydrogen is removed from the interstellar gas by star formation \\citep{searle72}. The stellar metallicity $Z_*$ is derived from $S(Z)/S_1 = (1 - f_1^{Z/Z_1})/(1 - f_1)$ and $Z_* = \\int^{Z_1}_0 ZdS(Z)/ \\int^{Z_1}_0 dS(Z)$, where $S(Z)$ is the total mass of stars born up to a time when the metal abundance reached to the value $Z$, and $S_1, Z_1, f_1$ denote $S(t_1), Z(t_1), f(t_1)$, respectively \\citep{pagel75}. In our case, $t_1$ is a time when the stellar metallicity reached to $Z_* = 2.5 Z_{\\sun}$. When the gas mass is reduced to $ f = 0.3$, we have $Z_* = 2.4 - 2.7 Z_{\\sun}$ and $Z_g = 4.5 - 5.1 Z_{\\sun} $. The inferred galaxy mass is $1.2 \\times$ the stellar mass at $t =0$, because the total mass ejected from stars at the events of SN explosions into the interstellar gas is 0.1 per unit stellar mass at $t =0$. The total gas mass is $2.3 \\times 10^{10} M_{\\sun}$. The gas-to-dust mass ratio is then $120 \\pm 73$, which is comparable to the value obtained in a $z = 1.3$ HyLIRG ($198 \\pm 53$) by \\citet{iono06}. The dust production rate is $0.24 \\pm 0.12 M_{\\sun}$ per SN. This is consistent with the numerical results of dust formation in SNe; for progenitor masses ranging 13 - 30 $M_{\\sun}$, 2 - 5\\% of the progenitor mass is locked into dust grains at SN explosions \\citep{nozawa03}, 20 - 100\\% of which is destroyed by processing through the collisions with the reverse shocks resulting from the interaction of SN ejecta and with the ambient medium \\citep{nozawa07}. \\subsection{Foreground dust screen} As demonstrated in Section 3.2, a foreground dust screen geometry is plausible for SST J1604+4304. This indicates that dust is depleted in the starburst site and the stellar spectra are reddened by the foreground dust screen which enclosed the starburst site. Let's suppose a spherical shell with a radius $l$ for the dust distribution, then we observe $f_{\\nu} = 4 \\tau_{\\nu}B_{\\nu}(T_{dust})\\pi(l/D)^2$ in the rest-frame, where the thickness of the shell is $\\tau_{\\nu}$ which is 0.0074 at rest 80\\micron\\ corresponding to $E(B-V)$ = 0.83 with $R_V$ = 4.05 \\citep{draine03}, $D$ is the distance from the observer to SST J1604+4304, and $T_{dust} = 32.5 - 35 K$. Converting this relation into the observer's frame, we obtain the radius $l$ = 4.5 - 5.5 kpc or 0.56 - 0.65\\arcsec. This is comparable to the optical size of SST J1604+4304 which is approximately 1.2\\arcsec $\\times$ 0.5\\arcsec in the $I_{814}$ band. This supports the view that the galaxy is surrounded by the shell. Further support is the fact that there are no significant color gradients across the galaxy in the $HST$ images. \\subsection{Evolutionary link to other galaxy populations} As discussed by \\citet{calzetti01}, starburst environments are rather inhospitable to dust; dust grains in the starburst site can be transported to a large distance in a relatively short time by radiation pressure(\\citealt{ferrara91}; \\citealt{venkatesan06}), as well dust grains that formed at SN explosions are processed and evolve in SN remnants \\citep{nozawa07} - small size grains are quickly destroyed in SN remnants by sputtering. Thus, a cavity-shell structure is a natural geometry for the dust distribution in star-forming galaxies; the starburst site is inside the cavity where dust is depleted, and the opaque dust shell is surrounding the cavity. The shell is observed as a foreground screen. Such foreground screens were found in local starburst galaxies (\\citealt{calzetti94}; \\citealt{meurer95}). In the foreground screen of SST J1604+4304, the dust would be unevenly distributed; we are only looking at part of the starburst site at UV and optical wavelengths through relatively transparent holes with $E(B-V)$ = 0.8, and the other part is completely obscured at these wavelengths. This explains that $L_{ir}$ is two times greater than the stellar luminosity derived from the broad-band SED analysis. \\citet{venkatesan06} and \\citet{nozawa07} suggest that dust created in the first SN explosions can be driven through the interior of the SN remnants and accumulated in the SN shells, where second-generation stars may form in compressed cooling gases. Hence, galaxies may be observed as dust-free objects at the very beginning. When much dust formed in SN ejecta, the starburst site would become completely obscured by dust at UV to near-infrared wavelengths. This stage would correspond to extremely dusty objects like DOGs and SMGs. Then, the galactic winds would turned on. \\citet{heckman90} discussed that the galactic winds or superwinds frequently observed in starbursts and ULIRGs will sweep out any diffuse interstellar matter from the starburst site. Such galactic winds will turned on when the thermal energy of the gas heated by SN explosions exceeds the gravitationally binding energy of the gas. According to the models for chemical evolution of elliptical galaxies by \\citet{arimoto87}, galactic winds turn on later in more massive galaxies. The onset of a galactic wind is 350 Myr for a $10^{11} M_{\\sun}$ galaxy, and the metallicity increases up to more than the solar value. These predictions seem to agree well with the age and the metallicity observed in SST J1604+4304, in which the galactic wind would just turn on and dust grains in the opaque shell are pushed and moved outward, creating partially transparent holes in the shell. This picture is consistent with the fact that dusty objects are deemed to be more abundant in massive galaxies than in less massive galaxies, although mid- and far-infrared observations are strongly biased to IR-luminous objects. This observational trend is expected, because galactic winds turn on later and thus the dust shell becomes more opaque for more massive galaxies, resulting in longer time of the dust-obscured stage. This may be the reason why DOGs and SMGs are so infrared-luminous and massive. \\subsection{Impact by an obscured AGN } What would happen if the missing 1/2 of the bolometric luminosity comes from an AGN? In this case, the stellar mass ranging 0.1 - 8 $M_{\\sun}$ is reduced to one half of that for the pure star formation. The number of SNe is also reduced by the same amount, while the dust mass does not change. Thus, the dust production rate is increased by a factor of 2, i.e., $0.48 \\pm 0.24 M_{\\sun}$ per SN. This is within a range consistent with the model predictions(\\citealt{nozawa03}; \\citealt{nozawa07}). The metallicity is the same as derived for pure star formation, because the SN ejecta mass is also reduced by a factor of 2. If the AGN shines at the Eddington rate, the mass of the AGN is $2.5 \\times 10^7 M_{\\sun}$." }, "0910/0910.0332_arXiv.txt": { "abstract": "Transverse stratification is a common intrinsic feature of astrophysical jets. There is growing evidence that jets in radio galaxies consist of a fast low density outflow at the jet axis, surrounded by a slower, denser, extended jet. The inner and outer jet components then have a different origin and launching mechanism, making their effective inertia, magnetization, associated energy flux and angular momentum content different as well. Their interface will develop differential rotation, where disruptions may occur. We here investigate the stability of rotating, two-component relativistic outflows typical for jets in radio galaxies. For this purpose, we parametrically explore the long term evolution of a transverse cross-section of radially stratified jets numerically, extending our previous study where a single, purely hydrodynamic evolution was considered. We include cases with poloidally magnetized jet components, covering hydro and magnetohydrodynamic models. With grid-adaptive relativistic magnetohydrodynamic simulations, augmented with approximate linear stability analysis, we revisit the interaction between the two jet components. We study the influence of dynamically important poloidal magnetic fields, with varying contributions of the inner component jet to the total kinetic energy flux of the jet, on their non-linear azimuthal stability. We demonstrate that two-component jets with high kinetic energy flux, and an inner jet effective inertia which is higher than the outer jet effective inertia are subject to the development of a relativistically enhanced, rotation-induced Rayleigh-Taylor type instability. This instability plays a major role in decelerating the inner jet and the overall jet decollimation. This novel deceleration scenario can partly explain the radio source dichotomy, relating it directly to the efficiency of the central engine in launching the inner jet component. The FRII/FRI transition could then occur when the relative kinetic energy flux of the inner to the outer jet grows beyond a certain treshold. ", "introduction": "There is strong observational and theoretical evidence that magnetic fields play a crucial role in the acceleration and the collimation of extragalactic jets. Most Active Galactic Nuclei (AGN) jet formation scenarios involve magnetic fields threading a rotating black hole (in the ergosphere) and its accretion disk, thereby removing from them angular momentum, allowing the central black hole to accrete. At least close to the jet launching region, the jet rotation profile persists, reflecting its (general relativistic and/or magneto-rotational) origin. Thus both ingredients, magnetic fields and rotation, are very important in jet formation, as well as in jet propagation and stability. Moreover, detailed astrophysical jet observations point out that relativistic jets are structured, in a direction perpendicular to the jet axis, typically consisting of a fast spine and slower outer flow. In the case of AGNs, this jet structuring plays an important role in explaining the morphology of the jet high energy radiation \\citep{Ghisellinietal05,Hardcastle06, Jesteretal06,Jesteretal07,Siemiginowskaetal07,Kataokaetal08}, with sometimes clear evidence for a very fast, light inner jet and a heavy slow outer outflow \\citep{Girolettietal04}. Furthermore, observations of the TeV BL Lacertae objects show brightenings and rapid variability in their TeV emission. This variation in their high energy emission implies high Lorentz factor flows occuring at smaller scale, suggesting ultra-relativistic bulk motion of the (inner) jet. At the same time, complementary (radio) observations with VLBI of the pc-scale jet structure indicate a broad, `slowly' (albeit relativistic) moving outflow. In combination, this clearly suggests the presence of a two component jet morphology \\citep{Ghisellinietal05}. Two-component jet structure has also been proposed in more theoretical work, addressing the physics of jet launching, collimation and propagation mechanisms \\citep{BogovalovTsinganos01, Soletal89, Meier03}. While our jet dynamics computations will be representative for AGN jet conditions, radially structured jet flows are now known to exist in virtually all astrophysical jet outflows. Transversely structured, ultra-relativistic jet-like outflow has been proposed in the context of Gamma Ray Bursts \\citep{Racusinetal08}, to explain the break observed in their afterglow light curve. In the case of stellar outflows, recent observations of some T Tauri jets \\citep{Bacciottietal00, Guntheretal09} also suggest a fast inner outflow bounded by a slow outer outflow. In these young stellar objects, a clear signature of jet rotation around the symmetry axis was detected \\citep{Bacciottietal02, Woitasetal05,Coffeyetal04,Coffeyetal07}, fully supporting scenarios of magneto-centrifugal jet launch and acceleration. Theoretical models of two-component jets in classical T Tauri \\citep{BogovalovTsinganos01,Melianietal06,Cranmer08,Fendt09} then postulate that the inner outflow is turbulent and pressure driven, associated with the young star wind. The inner jet then has a small opening angle, as it is collimated by the outer jet, which is in turn magneto-centrifugally driven from the surrounding disk. The outer disk wind then carries most of the mass loss in the jet. Various authors \\citep{Melianietal06, Fendt09} have demonstrated using axisymmetric MHD simulations that the outer outflow is self collimated by its intrinsic magnetic field, and that the turbulent inner outflow gets collimated by the outer jet. Furthermore, \\cite{Matsakos08} investigated the topological stability of two-component outflows for young stellar objects, performing extensive numerical simulations to determine whether analytic self-similar models demonstrate robustness in axisymmetric conditions. Also for relativistic jet simulations, axisymmetry assumptions are often adopted, excluding the development of all non-axisymmetric perturbations. These can address details of how helical field configurations (naturally expected from magneto-centrifugal launch mechanisms) effectively may transport their helicity down the jet beam \\citep{Keppensetal08}, with magnetically aided reacceleration by field compression across internal cross-shocks. While \\cite{Keppensetal08} concentrated on kinetic energy dominated jets, initially Poynting flux dominated jets were simulated by \\cite{Komissarovetal07} in axisymmetric relativistic MHD, finding that the transition to a matter-dominated jet regime occurs very close to the central engine (within $0.01 {\\rm pc}$). Our model computations will therefore assume kinetic energy dominated jets. As far as the magnetic field topology is concerned, we will restrict ourselves in this paper to purely poloidally magnetized jet components. Our two-component jet model determining our initial conditions can actually allow for helical fields, as explained in Sect.~\\ref{s-model} (we include this more general case here, for future reference in follow-up studies). As indicated before, during the first acceleration phase of AGN jets, magneto-centrifugal mechanisms play an important role and a helical or even strongly toroidal magnetic field is likely produced \\citep{Fendt97,Melianietal06a, McKinney&Blandford09,Komissarovetal07}. \\cite{McKinney&Blandford09} present 3D general relativistic MHD simulations for rapidly rotating black holes producing jets with strong toroidal fields. They find a prominent role of the accreted magnetic field geometry for achieving `stable' jets. Helical or strongly toroidal field topologies can be subject to current-driven kink instabilities \\citep{Begelman98}, with $m=1$ toroidal modes that helically displace the jet axis. This requires full 3D numerical simulations, such as performed by \\cite{BatyKep02} in non-relativistic MHD, or addressed by \\cite{Mizunoetal09} in relativistic MHD for a static force-free equilibrium. Dispersion relations for non-axisymmetric modes and $m=1$ kinks in particular for relativistic MHD were analysed by \\cite{Begelman98} for purely toroidal fields, and electromagnetically dominated, force-free jets were analysed spectrally by \\cite{Istomin&Pariev96} and more recently by~\\cite{Narayanetal09}. \\begin{figure}[h] \\begin{center} {\\resizebox{0.95\\columnwidth}{6cm}{\\includegraphics{Fig1.jpg}}} \\end{center} \\caption{3D schematic view of the overall AGN disk-jet configuration (indicating the accretion disk and the two-component jet). We model the jet evolution in the transverse plane.}\\label{Fig1} \\end{figure} In our work, we will restrict attention to 2.5D scenarios with the somewhat unusual assumption of translational invariance along the jet axis. The overall configuration is schematically indicated in Fig.~\\ref{Fig1}, and we simulate a transverse cross-section of the jet at sufficient distance from the two-component jet source, where all three velocity components (axial, azimuthal and radial) are included, but their variation along the jet axis is ignored. Our aim is to investigate all non-axisymmetric instabilities, primarily induced by the (sheared) rotation. This approximation is valid because in the poloidal direction, the flow is supersonic with high Lorentz factor, and then the growth rate of poloidal instabilities is expected to be low. On the other hand, the rotation is subsonic, facilitating the growth of toroidal instabilities. We then address jet stability in a cross section of a rotating two-component jet, initially collimated by thermal pressure and/or poloidal magnetic field. Consecutive snapshots of the cross-sectional evolution can be interpreted as mimicking the jet flow conditions at increasing distance from the source. Note that this particular assumption allows us to follow both axisymmetric and all non-axisymmetric (including $m=1$) mode development, but does exclude helical mode axis displacement typical for kink modes. Our model therefore mimics jet evolution adequately as long as the radial axis displacement is smaller than the axial wavelength associated with possible $m=1$ kinks. Our assumption does also neglect conical jet expansion, assuming cylindrical propagation. This is justified given the low observed values for jet opening angles. Since we defer the study of toroidal and helical magnetic field configurations in 2.5D and 3D to later work, we start off with numerically investigating the influence of a purely poloidal magnetic field on the stability of rotating, two-component relativistic jets. Our work complements the studies looking into kink development, by putting the emphasis on the azimuthal variation and on the effect of the two-component jet stratification. As far as a purely poloidal magnetic field topology is concerned, the work by \\cite{Spruitetal97} suggests that jets should be collimated by poloidal magnetic field pressure, rather than by toroidal magnetic field, as toroidal jet magnetic fields can introduce kink instability (but only slow mode growth was found for force-free jets by~\\cite{Narayanetal09}). To justify purely toroidal fields, we can argue that the toroidal magnetic field in the jet acceleration phase can be gradually dissipated involving reconnection, a mechanism which in turn contributes to (axial) jet acceleration \\citep{Spruit08, Melianietal06}. Indeed, during the acceleration phase a fraction of the Poynting flux (angular momentum) carried by the magnetic field is converted to kinetic energy by internal dissipation/reconnection of the toroidal magnetic field~\\citep{Spruit08, Melianietal06}. In accord with this mechanism, we model the region where the eventual jet rotation is low. In this paper, we analyse five cases in detail, to determine the effects of differing poloidal magnetic field configurations in the two-component structure on its long-term stability, and on the overall deceleration of the jet as it propagates away from the central source regions. The role of poloidal magnetic field is in these cases most prominent in its added effect on total pressure and effective fluid inertia, and this is shown to play a prominent role in the two-component jet stability. ", "conclusions": "We examined five configurations of magnetized two-component jets. All share the same density ratio between inner and outer component and identical rotation profiles. The magnetic and thermal pressure configuration in each model differs, though. This in turn translates to different distributions of the (total, fixed) kinetic energy flux over the inner and outer jet component. In fact, in case (A) the kinetic energy flux in the inner component is about $10\\%$, while it is $0.7\\%$ for case (B1) and $0.5\\%$ in case (B2), and around $38\\%$ for cases (C) and (D). For the outer component, we thus have $90\\%$ in case (A), $99.3\\%$ in case (B1) and $99.5\\%$ in case (B2), and only $62\\%$ for cases (C) and (D). The most important difference between the two model categories is then: cases (A, C, D) have an inner jet component with higher inertia $\\gamma^2\\,\\rho\\,h+B_{\\rm z}^2$ than their outer jet component, while cases (B1) and (B2) have an inner jet component inertia which ends up lower than in the outer jet component. From the detailed analysis of the simulations, as well as from the approximate stability analysis, this criterion distinguishes between cases where relativistically enhanced Rayleigh-Taylor modes ultimately dominate the evolution, leading to complete mixing of both components and inner jet deceleration. This is quantified most clearly by showing the time evolution of the mean Lorentz factor over the inner jet region for all cases in Fig.~\\ref{Lorentzfactor_evolution}. This requires a clear criterion to distinguish inner versus outer jets in the turbulent evolutions. In cases (A), (C) and (D), we locate the outer jet component as having a Lorentz factor $2.5<\\gamma<3.5$ and effective polytropic index $\\Gamma_{\\rm eff}>3/2$. The inner component is the region defined by a Lorentz factor $\\gamma\\ge 3.5$ and effective polytropic index $\\Gamma_{\\rm eff}\\le 3/2$. In the cases (B1) and (B2), the effective polytropic index in the inner and outer jet can be locally of the same order there, hence we use that the inner jet is magnetized. In the cases (B1) and (B2), the inner component jet and shear region are not totally mixed during the evolution, since the inner component is compressed and has higher magnetization and Lorentz factor. Thus to distinguish the inner component jet we put a condition on magnetic field strength $B_{\\rm z}>B_{\\rm z, initial}/2$, where $B_{\\rm z, initial}$ is the magnetic strength assumed initially. Under these precise quantifications of inner/outer jet regions, Fig.~\\ref{Lorentzfactor_evolution} demonstrates clearly how stable cases (for the relativistically enhanced Rayleigh-Taylor modes) remain at high speed, while unstable cases decelerate. Using the same means to distinguish inner versus outer jet regions at all times, we can quantify the inner jet radius for all cases, as well as the total jet radius for all cases. These are shown in Figs.~\\ref{Inradius_evolution}-\\ref{Radius_evolution}, and quantify the decollimation effects discussed in Section~\\ref{Simulations}. During the entire evolutions, the toroidal and radial speeds remain weak as we have typically a maximal $V_{R}<0.01$ and $V_{\\varphi}<0.04$. This means that the contribution of the laboratory frame charge separation force $\\rho_{e}{\\vec{E}}$ to the Lorentz force is negligible at all times. Under these conditions, in both radial and toroidal directions, the contribution of magnetic energy to fluid inertia is $1/\\gamma^2$ weaker than the contribution of the magnetic pressure to the total pressure. This explains why cases (B1) and (B2) are then more stable than the other cases. Despite the fact that case (B2) has similar axial total pressure than other cases (A, C, D), the low contribution of thermal energy to total pressure makes the effective inertia ratio between the inner and outer jet low. This two-component jet is then stable against the relativistically enhanced Rayleigh-Taylor instability. Extraction of angular momentum and energy from the inner jet to the shear shell is less efficient in cases (B1) and (B2) than in all other cases. As a result, in cases (B1) and (B2) the inner relativistic jet with high Lorentz factor $\\gamma_{\\rm in}\\sim 20$ persists. In the other cases, the kinetic energy flux in the inner jet was initially relatively high, making them unstable and leading to deceleration to Lorentz factors around $8$. We investigated the stability of two-component jets beyond the launching region, where both components are rotating differently and a clear two-component structure exists. We initialized this model in accord with magneto-centrifugal models for jet generation, also using the observed analogy between radio source jets and two-component jets in young stellar objects, where the rotation within the jet can actually be observed. We performed five very high resolution simulations of magnetized two-component jets with various magnetization and kinetic energy flux stratifications. The two-component jets with a low inner kinetic energy flux contribution are more stable and remain relativistic for long distances, whereas jets with a highly contributing inner jet to the total jet kinetic energy flux, are subject to a relativistic Rayleigh-Taylor type instability. This instability turns out to be very efficient to decollimate and decelerate the inner jet. Jets that are subject to this instability become turbulent after propagating for a distance of about $30 {\\rm pc}$. This new result on two-component jet models is important because it can explain the classification of radio sources in Fanaroff-Riley I/II categories according the energy stratification of the inner jet. This ultimately relates to the jet launch region and the properties of the inner accretion disk. In fact, by analogy between the FRI/FRII classification and the results of our model, an FRI jet would correspond to a two-component jet with a high energy flux contribution from the inner jet, whereas the FRII jet corresponds to relatively low energy fluxes in the inner jet. The model we propose here to explain the FRI/FRII dichotomy is different from the model we proposed earlier~\\citep{Melianietal08} where the transition occurs due to external density stratification. That model explains the group of peculiar ``HYbrid MOrphology Radio Source\" (HYMORS) \\citep{GopalKrishna&Wiita00} which appear to have an FR II type on one side and an FR I type diffuse radio lobe on the other side of the active nucleus. Since the launch conditions on each side are presumably similar in these kind of radio sources, the different Fanaroff-Riley morphologies on either side must be attributed to the change in the properties of the ambient media, as shown convincingly in~\\cite{Melianietal08}. The results of the present paper nicely complement these earlier findings with a quantifiable role of the central engine contribution. We currently continue this study of the interaction between two component jets in full 3D. We thereby intend to explore the relative influence of azimuthal versus longitudinal instabilities for realistic multi-component jets. Another extension is to allow for aximuthal magnetic fields, in accord with the initial profiles as given generally in this paper. It then rremains to be shown that (1) the newly discovered instability persists in 3D hydro and magnetohydrodynamic configurations, where the potential role of axial mode development is incorporated, and introduces helical jet axis displacements; (2) how the instability gets modified (stabilized or destabilized) by the inclusion of toroidal field components, first in 2.5D neglecting helical axis displacements, and consecutively in 3D, where also current-driven kinks may occur." }, "0910/0910.5721_arXiv.txt": { "abstract": "One of the primary means of determining the spin $a$ of an astrophysical black hole is by actually measuring the inner radius $r_\\mathrm{in}$ of a surrounding accretion disk and using that to infer $a$. By comparing a number of different estimates of $r_\\mathrm{in}$ from simulations of tilted accretion disks with differing black-hole spins, we show that such a procedure can give quite wrong answers. Over the range $0 \\le a/M \\le 0.9$, we find that, for moderately thick disks ($H/r \\sim 0.2$) with modest tilt ($15^\\circ$), $r_\\mathrm{in}$ is nearly independent of spin. This result is likely dependent on tilt, such that for larger tilts, it may even be that $r_\\mathrm{in}$ would increase with increasing spin. In the opposite limit, we confirm through numerical simulations of untilted disks that, in the limit of zero tilt, $r_\\mathrm{in}$ recovers approximately the expected dependence on $a$. ", "introduction": "\\label{sec:intro} Classical, astrophysical black holes have only two defining properties: mass and angular momentum (or spin). As in other astrophysical contexts, the mass of a black hole can straightforwardly be determined by the application of Kepler's Third Law, provided the black hole is orbited by another observable object and the distance to the system is reasonably well known (at least to within a factor of $\\sin i$, where $i$ is the inclination of the binary orbit relative to the observer's line-of-sight). The spin $a=Jc/GM$, on the other hand, is a very different matter. So far, no direct measure of spin has been proposed, leaving only indirect means of inferring it. A commonly used method is based on making some measure of the ``inner radius\" of the accretion disk. In general relativity, stable circular orbits are not permitted all the way down to the ``surface'' of the black hole, but are instead restricted to lie outside the marginally stable limit $r_\\mathrm{ms}$. Beginning with \\citet{nov73}, it has commonly been assumed that the accretion disk will observe a similar restriction and truncate at $r_\\mathrm{ms}$. Because $r_\\mathrm{ms}$ has a well-known monotonic dependence on the spin of the black hole \\citep{bar72}, its association with an observed inner radius of the accretion disk is then used to infer the spin of the black hole. The observation of the inner radius can come either from modeling of the continuum emission in the thermally dominant state \\citep[e.g.][]{sha06,dav06} or from measuring relativistically broadened reflection features \\citep[e.g.][]{wil01}. However, it has been shown that there are many difficulties with such procedures. First, one has many choices for how to define the effective inner radius $r_\\mathrm{in}$ of the accretion disk, and some of these can vary considerably (factors of 2-3) from $r_\\mathrm{ms}$ \\citep{kro02}. Even if one restricts oneself to such observationally relevant measures as the ``radiation edge'' (the innermost radius from which significant luminosity emerges) or ``reflection edge'' (the smallest radius capable of producing significant X-ray reflection features), significant deviations may be observed \\citep{kro02, bec08}. For instance, \\citet{ago00} showed that significant emission can emerge from inside $r_\\mathrm{ms}$, thus precluding a direct correlation with $r_\\mathrm{in}$. Similarly \\citet{rey97} showed that the reflection edge is not necessarily tied to $r_\\mathrm{ms}$. Nevertheless, a strong result for untilted disks (disks that lie approximately in the symmetry plane of their black hole spacetimes and have their angular momenta aligned with the spins of their black holes) is that $r_\\mathrm{in}$ should gradually decrease with increasing black hole spin, i.e. as $a/M\\rightarrow 1$. Therefore, it might at least be hoped that a study of $r_\\mathrm{in}$ for a large ensemble of accreting black holes might yield some information about the range and distribution of black hole spins represented in the ensemble. However, we show in this Letter that there is a further complicating factor. Based on numerical simulations, we show that the effective inner radius of tilted disks (disks that do not have their angular momenta aligned with the spins of their black holes) yield very different results for $r_\\mathrm{in}$ than their untilted counterparts for all dynamical measures we have considered. A disk could be tilted for many reasons. In stellar mass binaries, the orientation of the outer disk is fixed by the binary orbit, whereas the orientation of the black hole is determined by how it became part of the system. If the black hole formed from a member of a preexisting binary through a supernova, then the black hole could be tilted if the explosion were asymmetric. If the black hole joined the binary through multi-body interactions, such as binary capture or replacement, then there would have been no preexisting symmetry, so the resulting system would nearly always harbor a tilted black hole. This same argument can be extended to AGN in which merger events reorient the central black hole or its fuel supply and result in repeated tilted configurations. Tilted disks are subject to additional torques due to differential Lense-Thirring precession \\citep{bar75}. Our results indicate that for moderately thick ($H/r \\sim 0.2$), tilted ($15^\\circ$) disks, $r_\\mathrm{in}$ is independent of spin, or, in some measures, actually {\\em increases} with increasing black hole spin. These results could have important consequences for observational efforts to constrain black hole spins. ", "conclusions": "\\label{sec:discussion} We have explored four possible measures of the effective inner radius of simulated black-hole accretion disks, based on surface density, inflow and turbulent velocities, and magnetic field structure. With all four measures, we found two consistent trends: 1) for untilted simulations $r_\\mathrm{in}$ closely follows the analytic form of $r_\\mathrm{ms}$; whereas 2) for $15^\\circ$ tilted simulations $r_\\mathrm{in}$ shows a flat or even increasing trend as $a/M\\rightarrow 1$. The fact that Figures \\ref{fig:density} - \\ref{fig:alpha} all show remarkable similarities suggests that our results are not due to a poor choice of definition for $r_\\mathrm{in}$. Also, although we have ignored the dependence of $r_\\mathrm{ms}$ on inclination in making our plots, Figure 5 of \\citet{fra07} shows that the deviation would be quite small for $i = 15^\\circ$, as would be appropriate for this work. We can get some estimate of the uncertainties associated with our results by comparing data obtained from the two 515H simulations, one done on the spherical-polar grid and one on the cubed-sphere. For the first three measures of $r_\\mathrm{in}$, where we were able to get estimates using both simulations, they agree remarkably well. Furthermore, for most of the measures the trend lines fit the data well, suggesting small uncertainties. We conclude that the discrepancy of $r_\\mathrm{in}$ between tilted and untilted simulations is a robust result, and must, therefore, be rooted in some physical process. The most likely agent of such a process would seem to be the standing shocks aligned along the line of nodes between the black hole symmetry plane and disk midplane \\citep{fra08}. These features are not present in simulations of untilted disks. Such non-axisymmetric shocks can be quite efficient at extracting angular momentum from the gas flow, which would reduce the effect of the black hole spin on $r_\\mathrm{in}$. In this work we have only considered two values of initial tilt: $\\beta_i = 0$ and $15^\\circ$. It stands to reason, however, that as $\\beta_i \\rightarrow 0$ the inner radius would approach the untilted values obtained in this work. On the other hand, for $\\beta_i > 15^\\circ$, it may well be that the discrepancies in $r_\\mathrm{in}$ would be even larger. In such a case, $r_\\mathrm{in}$ might truly increase with increasing spin. The results in this Letter then imply that, at least for moderately thick accretion disks ($H/r \\gtrsim 0.1$), measurements of $r_\\mathrm{in}$ are {\\em not} reliable predictors of $a$, unless one can independently confirm that the disk is not tilted. This could pose a problem for recent attempts to determine spin using so-called ``Hard'' state observations \\citep{mil06, rei08}, where the flow is generally taken to be thick. For very thin disks ($H/r \\lesssim 0.01$), the effect of tilt is expected to be much different. Such disks are expected to be subject to the Bardeen-Petterson effect \\citep{bar75}, where the midplane of the disk at small radii would be aligned with the symmetry plane of the black hole through the competing action of Lense-Thirring precession and viscous diffusion, while possibly leaving an outer disk that is tilted at large radii. Since the inner disk would be aligned with the symmetry plane of the black hole, the effective inner radius should be the same as for an untilted disk. This would apply for disks in the ``Soft'' or ``Thermally-Dominant'' state, with luminosities well below the Eddington limit, which are likely to be thin. Attempts to estimate the black hole spin by modeling the disk in this state may be unaffected by the cautions raised in this Letter. However, for sources with disk luminosities near or above the Eddington limit \\citep[e.g.][]{mid06}, the disk should again be thick, and the effect discussed here will apply. Even if the disk is thin, we reiterate that it is still important to independently fit for the inclination of the inner disk, since this may not be the same as the binary inclination. Other methods of determining black-hole spin, such as using quasi-periodic oscillation (QPO) frequencies, may not be affected by our results. For instance, QPO models based on resonant frequencies attributable to the black hole spacetime itself \\citep[e.g.][]{abr01} would likely not be strongly affected. In fact, the weak dependence of $r_\\mathrm{in}$ on $a$ detailed in this Letter actually helps one model of low-frequency QPOs based on the precession of a radially extended thick disk, by ensuring that the maximum frequency would not exceed $\\sim 10$ Hz regardless of black hole spin, as observed \\citep{ing09}. However, to date, no consensus model for QPOs has emerged, leaving us with few options for unambiguously determining black-hole spin. To end, we emphasize that it would be wrong to conclude from this paper that all tilted black hole accretion disks should appear to have $r_\\mathrm{in} \\approx 6 M$. Remember, the normalizations used in each of our plots were chosen simply to give values for the untilted disks that were reasonably similar to the values for $r_\\mathrm{ms}$. There is no other physical motivation for choosing those normalizations, and a different choice would have led to different numerical values for $r_\\mathrm{in}$. Further, we can not really say how the dynamically determined $r_\\mathrm{in}$ in this work would compare to values of $r_\\mathrm{in}$ measured from continuum modeling or iron-line fitting without doing more work. Specifically, we would need to generate synthetic disk images from the simulation data, and measure a radiation or reflection edge. Such work is already underway using the relativistic ray tracing code of \\citet{dex09}. Even then, there is the problem that the original simulations did not properly account for all of the important physical processes occurring in the disk, notably heating due to dissipation of turbulent energy and radiative cooling. That will have to await future simulations. What {\\em can} be concluded from this paper is that all tilted disks (with a tilt around $15^\\circ$) should appear to have the {\\em same} $r_\\mathrm{in}$, independent of spin. Our results may further imply that measurement of a small value for $r_\\mathrm{in}$ (without defining small) would require a rapidly rotating black hole {\\em and} an untilted disk, whereas a large value of $r_\\mathrm{in}$ would {\\em not} necessarily mean that the black hole is spinning slowly; it could simply be tilted." }, "0910/0910.2337_arXiv.txt": { "abstract": "Accretion of dark energy onto black holes will take place when dark energy is not a cosmological constant. It has been proposed that the time evolution of the mass of the black holes in binary systems due to dark energy accretion could be detectable by gravitational radiation. This would make it possible to use observations of black hole binaries to measure local dark energy properties, e.g., to determine the sign of $1+w$ where $w$ is the dark energy equation of state. In this Letter we show that such measurements are unfeasible due to the low accretion rates. ", "introduction": "Experimental evidence shows that the universe currently undergoes an accelerated expansion driven by an unknown form of energy, dubbed dark energy. This dominant energy component can -- at least phenomenologically -- be described as a perfect fluid with density $\\rho_{de}$ and a (possibly time dependent) equation of state $w\\equiv p_{de}/\\left(\\rho_{de}c^{2}\\right)$, where $p_{de}$ is the dark energy pressure. Current measurements are consistent with the dark energy density being approximately 70\\,\\% of the critical density of the universe, $\\rho_{de}\\sim 10^{-29}\\,\\mbox{g cm}^{-3}$, and the equation of state being close to $w=-1$, i.e., the value corresponding to a cosmological constant. Since the density is very low and we do not expect dark energy to cluster to any large extent, all the information we have on dark energy is through its impact on the largest scales of the universe via the expansion rate of the universe. One of the foremost experimental and theoretical tasks in cosmology is to pin down the equation of state $w$ of the dark energy, specifically to constrain or detect any deviations from the cosmological constant value of $w=-1$. It has been proposed that dark energy can be detected also on smaller scales \\cite{mersini,wang}. Any such local measurements would constitute a powerful confirmation of the existence of dark energy. In \\citet{mersini}, the possibility to measure local properties of dark energy using gravitational radiation from binary systems of supermassive black holes is investigated. The basic idea is that accretion of dark energy by black holes will lead to a mass change when dark energy is not a cosmological constant. The mass accretion would affect the dynamics of the system and in turn alter the waveform of the gravitational radiation produced by the binaries during the inspiraling phase. Measuring this waveform, through LIGO or LISA, would then make it possible to constrain the values of $w$ and $\\rho_{de}$. Alternatively, the change in orbital period of the system could be measured using electromagnetic radiation, if present. In this Letter we claim that this effect is negligible and that this form of local measurement of $\\rho_{de}$ and $w$ is not possible. In Section 2 of the Letter, we compare the mass accretion due to the dark energy with that of the interstellar medium and dark matter. In Section 3 we look at the radial separation needed for dark energy accretion to dominate over the effect over gravitational radiation in a binary system of two supermassive black holes. In both cases we show that dark energy accretion cannot be considered a measurable effect, even for a binary system. In Section 4 we consider two concrete examples to validate this conclusion. ", "conclusions": "We have investigated the effect of dark energy accretion in binary supermassive black hole systems. When comparing the mass change due to accretion from the interstellar medium, dark matter and dark energy, we find that the effect of dark energy accretion for all practical purposes is negligible. We also compare the effects from dark energy accretion and the loss of energy through gravitational wave emission. At the separations between the binary constituents needed for dark energy accretion to dominate over gravitational radiation, and under the assumption that the rotation of dark energy in the vicinity of the binary system is not greatly exceeding that of the binary system, we find the radial change induced by the accretion to be too small to be observable. Thus, even in an idealized setting with no gas accretion, the effect of dark energy accretion during the inspiraling phase does not have a measurable impact on the dynamics of the system. Doing local measurements of the equation of state of dark energy, as proposed in \\citet{mersini}, can therefore unfortunately not be considered possible." }, "0910/0910.5818_arXiv.txt": { "abstract": "The radioscience experiment is one of the on board experiment of the Mercury ESA mission BepiColombo that will be launched in 2014. The goals of the experiment are to determine the gravity field of Mercury and its rotation state, to determine the orbit of Mercury, to constrain the possible theories of gravitation (for example by determining the post-Newtonian (PN) parameters), to provide the spacecraft position for geodesy experiments and to contribute to planetary ephemerides improvement. This is possible thanks to a new technology which allows to reach great accuracies in the observables range and range rate; it is well known that a similar level of accuracy requires studying a suitable model taking into account numerous relativistic effects. In this paper we deal with the modelling of the space-time coordinate transformations needed for the light-time computations and the numerical methods adopted to avoid rounding-off errors in such computations. ", "introduction": "\\label{intro} BepiColombo is an European Space Agency mission to be launched in 2014, with the goal of an in-depth exploration of the planet Mercury; it has been identified as one of the most challenging long-term planetary projects. Only two NASA missions had Mercury as target in the past, the Mariner 10, which flew by three times in 1974-5 and Messenger, which carried out its flybys on January and October 2008, September 2009 before it starts its year-long orbiter phase in March 2011. The BepiColombo mission is composed by two spacecraft to be put in orbit around Mercury. The radioscience experiment is one of the on board experiments, which would coordinate a gravimetry, a rotation and a relativity experiment, using a very accurate range and range rate tracking. These measurements will be performed by a full 5-way link (\\cite{iess01}) to the Mercury orbiter; by exploiting the frequency dependence of the refraction index, the differences between the Doppler measurements (done in Ka and X band) and the delay give information on the plasma content along the radiowave path. In this way most of the measurements errors introduced can be removed, improving of about two orders of magnitude with respect to the past technologies. The accuracies that can be achieved are $10$ cm in range and $3 \\times 10^{-4}$ cm/s in range rate. How we compute these observables? For example, a first approximation of the range could be given by the formula \\begin{equation} r=|{\\bf r}|=|({\\bf x}_{\\rm sat}+{\\bf x}_{\\rm M})-({\\bf x}_{\\rm EM}+ {\\bf x}_{\\rm E}+{\\bf x}_{\\rm ant})| \\,\\, , \\label{eq:new} \\end{equation} which models a very simple geometrical situation (see Figure~\\ref{fig:range}). The vector ${\\bf x}_{\\rm sat}$ is the mercurycentric position of the orbiter, the vector ${\\bf x}_{\\rm M}$ is the position of the center of mass of Mercury (M) in a reference system with origin at the Solar System Barycenter (SSB), the vector ${\\bf x}_{\\rm EM}$ is the position of the Earth-Moon center of mass in the same reference system, ${\\bf x}_{\\rm E}$ is the vector from the Earth-Moon Barycenter (EMB) to the center of mass of the Earth (E), the vector ${\\bf x}_{\\rm ant}$ is the position of the reference point of the ground antenna with respect to the center of mass of the Earth. \\begin{figure}[h] \\begin{center} \\includegraphics[width=6cm]{fig_range.eps} \\end{center} \\caption{\\footnotesize Geometric sketch of the vectors involved in the computation of the range. SSB is the Solar System Barycenter, M is the center of Mercury, EMB is the Earth-Moon Barycenter, E is the center of the Earth.} \\label{fig:range} \\end{figure} Using (\\ref{eq:new}) means to model the space as flat arena ($r$ is an Euclidean distance) and the time as absolute parameter. This is obviously not possible because it is clear that, beyond some threshold of accuracy, these quantities have to be formulated within the framework of Einstein's theory of gravity (general relativity theory, GRT). Moreover we have to take into account the different times at which the events have to be computed: the transmission of the signal at the transmit time ($t_t$), the signal at the Mercury orbiter at the time of bounce ($t_b$) and the reception of the signal at the receive time ($t_r$). Formula (\\ref{eq:new}) could be a good starting point to construct a correct relativistic formulation; with the word ``correct'' we do not mean all the possible relativistic effects, but the effects that can be measured by the experiment. This paper deals with the corrections to apply to this formula to obtain a consistent relativistic model for the computations of the observables and the practical implementation of such computations. In Section \\ref{sec:refsys} we discuss the relativistic four-dimensional reference systems used and the transformations adopted to make the sums in (\\ref{eq:new}) consistent; according to \\cite{soffel03}, with ``reference system'' we mean a purely mathematical construction, while a ``reference frame'' is a some physical realization of a reference system. The relativistic contribution to the time delay due to the Sun's gravitational field, the Shapiro effect, is described in Section \\ref{sec:shap}. Section \\ref{sec:lt} deals with the theoretical procedure to compute the light-time (range) and the Doppler shift (range rate). In Section~\\ref{sec:num} we discuss the practical implementation of the algorithms showing how we eliminate rounding-off problems. The equations of motion for the planets Mercury and Earth, including all the relativistic effects (and potential violations of GRT) required to the accuracy of the BepiColombo radioscience experiment have already been discussed in \\cite{mil09b}, thus this paper concentrates on the computation of the observables. ", "conclusions": "\\label{sec:conc} By combining the results of the previous paper (\\cite{mil09b}), and of this one, we have completed the task of showing that it is possible to build a consistent relativistic model of the dynamics and of the observations for a Mercury orbiter tracked from the Earth, at a level of accuracy and self-consistency compatible with the very demanding requirements of the BepiColombo radioscience experiment. In particular, in this paper we have given the algorithm definitions for the computation of the observables range and range rate, including the reference system effects and the Shapiro effect. We have shown which computation can be performed explicitly and which ones need to be obtained from an iterative procedure. We have also shown how to push these computations, when implemented in a realistic computer with rounding-off, to the needed accuracy level, even without the cumbersome usage of quadruple precision. The list of ``relativistic corrections'', assuming we can distinguish their effects separately, is long, and we have shown that many subtle effects are relevant to the required accuracy. However, in the end what is required is just to be fully consistent with a post-Newtonian formulation to some order, to be adjusted when necessary. Interestingly, the high accuracy of BepiColombo radio system may require implementation of the second post-Newtonian effects in range." }, "0910/0910.5735_arXiv.txt": { "abstract": "Current and future weak lensing surveys will rely on photometrically estimated redshifts of very large numbers of galaxies. In this paper, we address several different aspects of the demanding photo-z performance that will be required for future experiments, such as the proposed ESA Euclid mission. It is first shown that the proposed all-sky near-infrared photometry from Euclid, in combination with anticipated ground-based photometry (e.g. PanStarrs-2 or DES) should yield the required precision in individual photo-z of $ \\sigma_{z} (z) \\leq 0.05(1+z)$ at $ I_{AB} \\leq 24.5$. Simple a priori rejection schemes based on the photometry alone can be tuned to recognise objects with wildly discrepant photo-z and to reduce the outlier fraction to $ \\leq 0.25\\%$ with only modest loss of otherwise usable objects. Turning to the more challenging problem of determining the mean redshift $ \\langle z\\rangle$ of a set of galaxies to a precision of $ |\\Delta_{\\langle z \\rangle}|\\leq 0.002(1+z)$ we argue that, for many different reasons, this is best accomplished by relying on the photo-z themselves rather than on the direct measurement of $\\langle z\\rangle$ from spectroscopic redshifts of a representative subset of the galaxies, as has usually been envisaged. We present in an Appendix an analysis of the substantial difficulties in the latter approach that arise from the presence of large scale structure in spectroscopic survey fields. A simple adaptive scheme based on the statistical properties of the photo-z likelihood functions is shown to meet this stringent systematic requirement. We also examine the effect of an imprecise correction for Galactic extinction on the photo-z and the precision with which the Galactic extinction can be determined from the photometric data itself, for galaxies with or without spectroscopic redshifts. We also explore the effects of contamination by fainter over-lapping objects in photo-z determination. The overall conclusion of this work is that the acquisition of photometrically estimated redshifts with the precision required for Euclid, or other similar experiments, will be challenging but possible. ", "introduction": "Large scale mapping of the weak lensing shear field in three dimensions is emerging as a potentially very powerful cosmological probe \\citep{Peacock2006, DETF}. Weak lensing has the advantage of directly tracing the mass distribution, thereby bypassing much of the complex astrophysics of the baryon component that underpin most of the other probes and which may well dominate the systematic uncertainties in them. In contrast, the underlying physics of weak lensing is extremely simple, and the challenges are primarily on the observational side, particularly the accurate measurement of the weak lensing distortion and the estimation of distances to very large numbers of faint galaxies. A weak lensing survey of half the sky $ \\rm (20,000\\, deg^{2})$ to a depth of $ I_{AB} \\sim 24.5$ and with a PSF of $ \\sim$ 0.2 arcsec, forms a major part of the proposed ESA Euclid mission\\footnote{http://www.ias.u-psud.fr/imEuclid \\\\ http://sci.esa.int/science-e/www/area/index.cfm?fareaid=102}. Euclid had its origins in two proposals submitted for the first round of the ESA Cosmic Visions 2015-2025 competition, the DUNE imaging survey \\citep{DUNE2006} and the SPACE spectroscopic survey \\citep{SPACE2009}. The combination of the two surveys, plus the anticipated improved information on the Cosmic Microwave Background from Planck\\footnote{www.rssd.esa.int/Planck} offers dramatic improvements in our knowledge of the entire dark sector, including the definition of the dark matter power spectrum, the dark energy equation of state parameter $ w$, as well as much else. Application of weak lensing for cosmology requires at least a statistical knowledge of the distances, i.e. redshifts, of large numbers of individual galaxies. At $ I_{AB} < 24.5$, there are about 2.5 billion galaxies in the Euclid $ 2\\pi$ sr survey area and so, realistically, reliance must be made on photometrically estimated redshifts (hereafter \\textit{photo-z}). \\subsection{Required redshift precision for precision cosmology with weak lensing} Several papers have discussed the redshift precision that is needed for weak lensing analyses to enable the full potential of this approach to be exploited \\citep{Amara&Refrigier2007, Ma&HUetal2006, Abdalla2008}. In the lensing tomography approach \\citep{Hu_Tomography}, individual galaxies are binned into a number of redshift bins. The shear signal is extracted from the cross-correlation of the shape measurements of individual galaxies in different redshift bins. These correlated alignments then give information (with some distance weighting function) on the mass distribution between the observer and the nearer of the two redshift bins. Redshift information for the galaxies is required at two conceptually distinct steps: first, the construction of the redshift bins used for the cross-correlation analysis to extract the weak lensing signal and, second, the estimation of the mean redshift of the galaxies in a given bin, which is required to map the results onto cosmological distance and thereby extract the cosmological parameters. It is, of course, possible to do a similar correlation analysis with unbinned data \\citep{2005PhRvD..72b3516C, kitching2008}, but for our purposes this distinction is unimportant. The cross-correlation between different redshift bins is undertaken to exclude any galaxy pairs that may be physically associated, i.e. have the same distance. This is to avoid the possibility that physical processes operating around individual galaxies may produce an intrinsic alignment of the galaxies that may be mistaken for the coherent alignment produced by the weak lensing of the foreground mass distribution. The required accuracy of the individual photo-z for the bin-construction task is set by the need to exclude overlaps in the N(z) of individual bins (or the probability distribution for individual galaxies) and thereby remove physically close pairs \\citep{2002A&A...396..411K}. This typically sets a requirement on the precision of individual photo-z of about $ \\sigma_{z} = 0.05(1+z)$ \\citep{bridle&King2007}. There is a second type of intrinsic alignment effect \\citep{2004PhRvD..70f3526H}, whereby the shape of the further of a given galaxy pair may be affected, through lensing effects, by the shape of the matter distribution around the nearer galaxy, which is likely to be correlated with the visible shape of that galaxy, thereby again producing a correlated alignment of the two galaxies that is unfortunately nothing to do with the lensing signal from the common foreground. \\cite{Jochimi&Schneider2008},~\\cite{Joachimi&Schnider2009} have shown that it is possible to implement a nulling approach to eliminate this second intrinsic alignment signal, which again requires a priori knowledge of individual redshifts. Once the weak lensing signal is extracted, accurate knowledge of the redshift of the galaxies, as with any cosmological probe gives, amongst other parameters, information on the angular diameter distance $ D_{\\theta}(z)$. The sensitivity of $ D_{\\theta}(z)$ to the relevant cosmological parameters ($ \\Omega_{m}$, $ \\Omega_{\\Lambda}$, $w$ etc) gives the required accuracy in the mean redshifts that are required to achieve a given precision on the parameters. As an example, \\cite{Peacock2006} have shown that a precision of 1$ \\%$ in $w$ requires a typical precision in the mean redshift of about 0.2$ \\%$ in $\\langle z\\rangle$. The Euclid goal is a 2$ \\%$ precision in $w$ (independent of priors), requiring a precision of order 0.002(1+z) in $\\langle z\\rangle$. This simple approach is confirmed by extensive analysis of the Fisher matrices \\citep{Amara&Refrigier2007, Ma&HUetal2006}. It is generally the case that if the mean redshift of a bin is defined accurately enough, then the higher moments of N(z) within the bin will also have been sufficiently determined. Of course, systematic biases in $\\langle z\\rangle$ that vary smoothly with redshift are particularly troublesome as they will mimic the effect of changing the cosmological parameters. In summary, in order to reach the Euclid performance, we require a statistical (random) r.m.s. precision of order $0.05(1+z)$ per galaxy (for the correlation analysis), and a systematic precision in the mean z in each bin of order 0.002(1+z). These are both quite demanding requirements, and together with the shape measurement itself ($ \\delta \\gamma \\sim 3 \\times 10^{-4}$ - \\citep{GREAT08, 2008MNRAS.391..228A}), they represent one of the observational challenges that lie along the path to enabling precision cosmology with weak lensing. Fortunately, there are some mitigating features of weak lensing analysis. For instance, the analysis is robust (aside from root-n statistics) to the exclusion of individual galaxies, provided only that the exclusion is unrelated to their shapes. One is free therefore to reject galaxies that are likely to have poor photo-z provided that they can be recognized a priori, i.e. from the photometric data alone. \\subsection{Challenges for the spectroscopic calibration of N(z)} Given the stringent requirements on the systematic error in the mean redshift $\\langle z\\rangle$ of a particular bin, one approach \\citep{Abdalla2008} is to define the $N(z)$ and mean $\\langle z\\rangle$ through the acquisition of spectroscopic redshifts for a representative subset of the galaxies. This direct sampling approach is certainly the most conservative, but will be very challenging, in practice, for the following reasons. First, one clearly requires very large numbers of spectroscopic redshifts. If we have a total redshift interval of $ \\Delta z$, split into m bins, then the number of spectroscopic redshifts N will be of order: \\begin{equation} N\\, \\sim \\, m^{-1} (\\Delta z/ \\sigma_{\\langle z\\rangle})^{2} \\end{equation} This assumes that the photo-z are perfect, and that there are no outliers. This leads trivially \\citep{Amara&Refrigier2007} to a requirement for $ 10^{5}-10^{6}$ spectroscopic redshifts. Secondly, these spectroscopic redshifts must be fully representative of the underlying sample. Any biases in the sampling of the bin or, even harder to reliably quantify, the almost inevitable biases in the ability to secure a reliable spectroscopic redshift, must be dealt with via a possibly complex and inevitably somewhat uncertain weighting scheme \\citep{limaetal2008}. It should be noted that current routine spectroscopic surveys of typical faint galaxies do not even approach 100$ \\%$ completeness, even at brighter levels. One of the best to date is the zCOSMOS survey \\citep{Lilly2007} on relatively bright $ I_{AB} < 22.5$ galaxies which yields, at present, a 99$ \\%$ secure redshift for 95$ \\%$ of galaxies at its optimum $ 0.5 < z < 0.8$ \\citep{Lilly2009}. Most other surveys are significantly less complete. Even more invidious are the effects of large scale structure in the spectroscopic survey fields, often called cosmic variance. Our own semi-empirical analysis (see the Appendix) of the COSMOS mock catalogues \\citep{kitzbichler&White2007} shows that, in a given patch of sky, the N(z) at $ I_{AB} \\sim 24$ becomes dominated by cosmic variance as soon as a rather small number of galaxies have been observed. The precise number depends on the field of view of the spectrograph, but is typically about 20-100 for spectroscopic fields of order 0.02-1 square degrees, i.e. a sampling rate of only a few percent. This means that the spectroscopic survey must be split up over a very large number of independent fields and that to get $ 10^{6}$ redshifts that are Poisson variance dominated one must effectively cover the whole sky in a sparse sampled way. This is unlikely to be efficient with the large telescopes needed for such faint object spectroscopy. A similar concern comes from the effects of Galactic extinction and reddening, which are likely, even when corrected for, to require spectroscopic sampling across the full range of Galactic [b,l]. These difficulties prompt consideration of other approaches, and in particular, that of placing greater reliance on the photo-z themselves, not only to construct the bins, but also to define their $\\langle z\\rangle$ with small systematic error. \\subsection{Using photo-z for construction of N(z)} The performance of photometric redshifts is continually improving. For example, in the COSMOS field \\citep{2007ApJS..172....1S}, where we have very deep 30-band photometry from the ultraviolet(GALEX) to $ 5\\mu m$, several photo-z schemes now achieve a precision of $\\sigma_{z} \\sim 0.01(1+z)$, with an outlier fraction (in non-masked areas and defined as a redshift difference greater than $0.15(1+z_{spec})$ of $ <1 \\%$ \\citep{2009ApJ...690.1236I}), at $ I_{AB} < 22.5$ and $ 0.05 < z < 1.4$, where the photo-z can be checked with about over 10,000 spectroscopic redshifts from zCOSMOS. It should be noted in passing that these 30 bands represent a very inhomogeneous data set in terms of point spread function, etc., and so this impressive photo-z performance also demonstrates the feasibility of combining disparate data into homogeneous photometric catalogues. Of course, this outstanding performance in the COSMOS field is unlikely to be achieved over the whole sky for the foreseeable future because of the expense of the required multi-band photometry. Nevertheless, the demonstration of this performance in COSMOS suggests that we have not yet reached any fundamental limit to photo-z performance. There are a number of different approaches to photo-z estimation that can be broadly distinguished between template-matching and more purely empirical approaches, such as Artificial Neural Networks \\citep{ANN_Lahav}. These have complementary strengths. Template fitting is based on the observed limited dimensionality of galaxy spectral energy distributions plus an astrophysical knowledge of the effects that can modify them, e.g. the the redshift itself, and the effects of extinction in our own Galaxy or in the distant galaxy. The more empirical approaches in essence avoid any such assumptions, which is both a strength and a limitation. Although both approaches have passionate adherents, our own view is that both approaches can normally perform equally well and that both are normally limited by the quality of the available data. In practice both use elements of the other, e.g. in template fitting, the actual data can be used to adjust the templates and the photometric zero-points, somewhat blurring the distinction. Finally, it should be noted that both can produce a likelihood distribution in redshift space through the application of priors \\citep{Hyperz, ANN_Lahav, BPZ, EASY}. In this paper we will base our analysis on a template-fitting algorithm (ZEBRA,~\\citep{2008arXiv0801.3275F}), since we believe its strengths are well suited to the problem in hand. We also note that the impressive real-life performance in COSMOS described above was achieved with two independent template fitting codes (Le Phare, \\citep{2009ApJ...690.1236I} and ZEBRA \\citep{Feldmann2006}). This improving photo-z performance described above suggests that it may be possible to use the photo-z themselves to construct the N(z), and thus $ \\langle z\\rangle$ for each bin, providing that the systematic uncertainties can be kept below the required level of 0.002(1+z). Some spectroscopic calibration would of course still be required, but the focus of this would be on constructing and characterizing the photo-z algorithm, rather than on constructing the N(z) directly. This approach would have a number of potential advantages over that discussed by \\cite{Abdalla2008} and others, and summarized above. At the very least, the number of spectra needed may be substantially reduced, although it is unlikely that one would wish to rely entirely on the photo-z and eliminate the spectroscopic conformation completely. However, to characterize the uncertainties in the photo-z, defined by $ \\sigma_{z}$, to the required level ($ \\sigma_{\\langle z\\rangle}$) we would need of order \\begin{equation} N \\sim (\\sigma_{z}/\\sigma_{\\langle z \\rangle})^{2} \\end{equation} spectroscopic redshifts, which may be orders of magnitude or more smaller than that implied by equation 1 since $ \\sigma_{z} \\sim 0.05 \\Delta z$. More importantly, the requirements on completeness and sampling are substantially relaxed since the photo-z characterization is done on individual objects. As one example, it is relatively easy to simulate the degradation in photo-z performance with noisier photometric data, so the calibrating spectroscopic objects need not necessarily extend all the way down to the photometric limit. The cosmic variance problem in spectroscopic calibration is eliminated, and the uncertainties arising from Galactic reddening can also be substantially reduced. \\subsection{Subject of this paper} The aim of this paper is to explore the performance of a template fitting photo-z code as applied to simulated photometric data of the approximate quality that we may realistically expect for a $ 2\\pi$ sr \\textit{all-sky} ground and space survey within the next decade. Our emphasis is on both the per object performance and on the potential for recognising and correcting systematic biases, which must be done to a high level if the increased reliance on photo-z is to be possible. As described in more detail in Section 3, we will assume for definiteness a photometric data set that includes the three-band near infrared photometry that is planned for Euclid itself, plus 5-band grizy photometry similar to that which should be produced by the PanStarrs-1, -2 and -4 projects (hereafter PS-1, -2 and -4)\\footnote{http://pan-starrs.ifa.hawaii.edu}. Examining these three generations of ground-based survey probes a range of depths that can be compared against other future surveys such as DES\\footnote{http://www.darkenergysurvey.org} and LSST\\footnote{http://www.lsst.org}. We then explore the following four topics that potentially may limit the photo-z performance and their usefulness to construct N(z) and $\\langle z\\rangle$: \\begin{enumerate} \\item The photo-z performance on individual objects in terms of the r.m.s. scatter (and bias) between the true redshift and the maximum likelihood photo-z, with particular emphasis on how to a priori identify and reject the outliers (\\textit{catastrophic failures}) from their individual photo-z L(z) likelihood distributions. \\item The construction of N(z) for a given set of photo-z selected galaxies, using their photo-z alone, with an emphasis on how to modify their individual likelihood functions L(z) to yield the least biased estimate of N(z) and $\\langle z\\rangle$ for the ensemble. \\item The systematic biases that can enter into the photo-z from an incorrect assumption about the level of foreground Galactic reddening, and how well the photometric data themselves can be used to determine the foreground reddening, both for a set of galaxies at known redshifts, and for those without known redshifts. \\item The effects of the photometric superposition of two galaxies at different redshifts, leading to a mixed spectral energy distribution that may perturb the photo-z, with an emphasis on seeing what happens to the redshift likelihood distribution. An interesting question is whether such composite objects can be recognised photometrically as well as ``morphologically'' from the images. \\end{enumerate} Our approach is to try to isolate these problems and to explore each in turn with the aim of providing an \\textit{existence proof} that provides a plausible route to achieve the very high photo-z performance that is required for Euclid. In particular, we decided to construct the input photometric catalogues using exactly the same set of approximately ten thousand templates as we subsequently used in the ZEBRA photo-z code. This may strike some readers as being somewhat circular. However, this approach allows us to eliminate the choice of templates as a variable, or uncertainty, in our analysis. This is motivated by the exceptional performance (discussed above) that has already been achieved with the same templates coupled with the exquisite observation data in COSMOS. This strongly suggests to us that the choice of templates is unlikely to be the limiting factor with the degraded photometry that we can realistically expect to have over the whole sky within the timescale of a decade or so. Although focused on the Euclid cosmology mission, the ideas and results from this paper may be of interest in many other applications that involve photo-z. ", "conclusions": "In this work we have investigated a number of issues that could potentially limit the photo-z performance of deep all-sky surveys, and thereby impede the ambitious precision cosmology goals of survey programs such as the proposed ESA Euclid mission. In each case, we find that, while standard techniques do not get to the required accuracy, simple new approaches can be developed that appear to get to the required performance. Knowledge of the redshifts enters into weak-lensing analyses at two distinct steps: first, the selection of objects for shape cross-correlation, and second the accurate knowledge of the mean redshifts of these objects. For practical reasons, the first step will likely require the use of photo-z for the foreseeable future. A major motivation for the current work has been to develop techniques that rely on photo-z also for the second step, bypassing the need for very large and highly statistically complete spectroscopic surveys (e.g. \\citep{Abdalla2008}). This approach thereby avoids the substantial practical difficulties that will be encountered in such spectroscopic surveys from incompleteness and the effects of large scale structure. A separate Appendix explores the latter effects in some detail. The work is based entirely on simulated photometric catalogues that have been constructed to match the expected performance of three generic ground-based surveys, combined with the expected near-IR imaging photometry from Euclid. To construct these catalogues, we use the same set of 10,000 templates as used for the template-fitting photo-z program (ZEBRA). This possibly circular approach allows us to remove the choice of templates as a variable. We believe it is justified at the current time by the very impressive performance of template-fitting photo-z codes applied to the deep multi-band COSMOS photometry, which strongly suggest that the choice of templates will not be a limiting factor at the required level, although further refinement will be desirable. The analysis is conveniently summarized in terms of the two main requirements on photo-z for precision weak lensing, namely to obtain an r.m.s. precision per object of $\\leq 0.05(1+z)$ and a systematic bias on the mean redshift of a given set of galaxies of $\\leq 0.002(1+z)$. Our main conclusions may be summarized as follows: \\begin{center} \\begin{itemize} \\item To achieve an r.m.s. photo-z accuracy of $ \\sigma_{z}(z)/(1+z) \\sim 0.05 $ down to $I_{AB} \\leq 24.5$, we need the combination of ground-based photometry with the characteristics of Survey-B (similar to PanStarrs-2 or DES) and the deep all-sky near-infrared survey from Euclid itself. This performance also requires the implementation of an ``a priori'' rejection scheme (i.e. based on the photometry alone, without knowledge of the actual redshifts of any galaxies) that rejects 13$ \\%$ of the galaxies and reduces the fraction of $ 5\\sigma$ outliers to below $ f_{cat} < 0.25\\%$. There is a trade off between the rejection of outliers and the loss of ``innocent'' galaxies with usable photo-z. Deeper photometry improves both the statistical accuracy of the photo-z and reduces the wastage in eliminating the catastrophic failures, and the combination of survey-C with Euclid near-infrared photometry achieves $ \\sigma_{z}(z)/(1+z) \\sim 0.04 $ after 9$ \\%$ rejection. \\item A good way to determine the actual $N(z)$ of a set of galaxies in a given photo-z selected redshift bin is to sum their individual likelihood functions. We find that the $\\sum L(z)$ function represents well both the wings of the $N(z)$ and the remaining catastrophic failures. outliers. However, to reach the required performance on the mean of the redshifts $| \\Delta_{\\langle z\\rangle} | \\leq 0.002(1+z)$ with the Survey-B, or deeper survey-C, combination (together with Euclid infrared photometry), we had to implement a modification scheme on the individual $L(z)$. This is based on the spectroscopic measurement of redshifts for a rather small number of galaxies (less than 1000) with relaxed requirements on statistical completeness (and no dependence on Large Scale Structure in the spectroscopic survey fields). We then require that the distribution of the actual redshifts within the probability space that is defined by the individual $L(z)$ should be flat across the sample as a whole. This should be true of any set of galaxies, leading to relaxed requirements on sampling of the redshift survey. This approach is similar to the application of a Bayesian prior on the redshifts, but is performed in probability space. Although it cannot be rigourously justified, it is found to work well in practice. \\item We find that uncertainties in foreground Galactic reddening can have a serious effect in perturbing the photo-z, with a net sign that varies erratically with redshift. However, we also find that such errors in $A_{v}$ can be identified internally from the photometric data of galaxies, either with or without spectroscopic redshifts. This procedure works best for galaxies with relatively high S/N photometry $ I_{AB} \\leq 22$. The required number of galaxies suggests that a reddening map on the scales of 0.1 deg$^{2}$ can be internally constructed from the data on galaxies without known redshifts, or from a few hundred galaxies with spectroscopic redshifts. \\item We also explored the effect of blended objects. The photo-z of the composite spectral energy distribution is a good representation of the redshift of the brighter object as long as the magnitude difference is large, i.e. $ \\Delta I_{AB} > 2.$ When the galaxies are more similar in brightness, $ \\Delta I_{AB} < 2$ there is a wide-range of behaviour. In some cases, multi-modal likelihood functions appear, while in others there is a sharp transition from one redshift to another, sometimes with a local maximum at a third, completely spurious redshift. In still others, the likelihood function smoothly transitions between the two redshifts with a single peak at an intermediate redshift. Our conclusion is that composite objects with $ \\Delta I_{AB} < 2$ should be recognized morphologically from imaging data and removed from the photo-z analysis. \\end{itemize} \\end{center} The general conclusion of our study is that while reaching the photo-z performance required for weak-lensing surveys such as Euclid will not be trivial, the implementation of new techniques, coupled with internal calibration of e.g. foreground reddening from the photometric data itself, will allow the required performance to be attained. If more reliance is placed on the photo-z themselves, then this leads to a significant simplification of the otherwise challenging requirement for spectroscopic calibration of large scale photometric surveys." }, "0910/0910.5359_arXiv.txt": { "abstract": "The description of high-energy hadronic interactions plays an important role in the (astrophysical) interpretation of air shower data. The parameter space important for the development of air showers (energy and kinematical range) extends well beyond todays accelerator capabilities. Therefore, accurate measurements of air showers are used to constrain modern models to describe high-energy hadronic interactions. The results obtained are complementary to information gained at accelerators and add to our understanding of high-energy hadronic interactions. ", "introduction": "The understanding and modeling of extensive air showers (particle cascades in the atmosphere) brings together the particle physics and astroparticle physics communities. To strengthen the connections and the scientific exchange between those communities is very fruitful for both sides and yields complementary information on the understanding of high-energy hadronic interactions. When high-energy cosmic rays impinge onto the atmosphere they initiate cascades of secondary particles -- the extensive air showers. Observations of air showers are used to indirectly infer the properties of cosmic rays at energies exceeding $10^{14}$~eV. The interpretation of air shower data faces a twofold challenge: the (exact) mass composition of cosmic rays is not known at those energies and, additionally, the properties of high-energy interactions taking place in air showers are partly unknown. Direct measurements of cosmic rays (fully ionized atomic nuclei) at energies below $10^{14}$~eV indicate that they are mostly composed of elements from hydrogen (protons) up to iron \\cite{cospar06,behreview}. The abundance of heavier elements is significantly smaller. Hence, in the following, we assume that cosmic rays comprise elements from hydrogen to iron. We will focus on results from the KASCADE experiment \\cite{kascadenim}, one of the most advanced air shower detectors in the energy range around $10^{15}$~eV. It has a unique set-up which allows to measure simultaneously the electromagnetic, muonic, and hadronic shower components. This is in particular valuable to test the consistency of hadronic interaction models. Since about a decade \\cite{Horandel:1998br,wwtestjpg} systematic checks of interaction models are performed with air shower data and the most stringent constraints on interaction models, derived from air shower data have been obtained with KASCADE measurements. KASCADE consists of several detector systems. A $200 \\times 200$~m$^2$ array of 252 detector stations, equipped with scintillation counters, measures the electromagnetic and, below a lead/iron shielding, the muonic parts of air showers. An iron sampling calorimeter of $16 \\times 20$~m$^2$ area detects hadronic particles \\cite{kalonim}. It has been calibrated with a test beam at the SPS at CERN up to 350~GeV particle energy \\cite{kalocern}. For a detailed description of the reconstruction algorithms see \\rref{kascadelateral}. ", "conclusions": "" }, "0910/0910.4396_arXiv.txt": { "abstract": "We study spectral distortions of diffuse ultra-high energy (UHE) neutrino flavour fluxes resulting due to physics beyond the Standard Model (SM). Even large spectral differences between flavours at the source are massaged into a common shape at earth by SM oscillations, thus, any significant observed spectral differences are an indicator of new physics present in the oscillation probability during propagation. Lorentz symmetry violation (LV) and Neutrino decay are examples, and result in significant distortion of the fluxes and of the well-known bounds on them, which may allow UHE detectors to probe LV parameters, lifetimes and the mass hierarchy over a broad range. ", "introduction": "Introduction} The neutrino sky spans about twenty five orders of magnitude in energy, potentially offering the possibility of probing the universe at widely disparate energy scales. The high end ($10^{11}-10^{12} $GeV) of this remarkably broad band in energy is set by: a) GZK neutrinos \\cite{PhysRevLett.16.748,Zatsepin:1966jv}, which originate in the interactions of the highest energy cosmic rays with the cosmic microwave photon background and b) neutrinos from the most energetic astrophysical objects observed in the universe, \\textit{i.e.} active galactic nuclei (AGNs) and Gamma-Ray bursts (GRBs). Detection, still in the future, presents both considerable opportunities and formidable challenges. In particular, at the very highest energies ($10^5$ GeV and above) which are the focus of this paper, the tiny fluxes that arrive at earth require detectors that combine the capability to monitor very large detection volumes with innovative techniques (for reviews, see e.g. \\cite{Hoffman:2008yu} and \\cite{Halzen:2007sz}). Examples of such detectors are AMANDA \\cite{Baret:2008zz}, ICECUBE \\cite{Halzen:2009tz}, BAIKAL \\cite{Aynutdinov:2009zz}, ANTARES \\cite{Margiotta:2009zz}, RICE \\cite{1742-6596-81-1-012008} and ANITA \\cite{1742-6596-136-2-022052}. A compelling motivation for exploring UHE neutrino astronomy is the fact that the origin of cosmic rays (CR) beyond the ``knee\" ($10^6$ GeV) remains a mystery many decades after their discovery. Additionally, CR with energies in excess of $10^{11}$ GeV have been observed\\cite{Roth:2009zz,Abbasi:2007sv}, signalling the presence of astrophysical particle accelerators of unprecedentedly high energies. If protons as well as neutrons are accelerated at these sites in addition to electrons, standard particle physics predicts correlated fluxes of neutrons and neutrinos which escape from the confining magnetic field of the source, while protons and electrons stay trapped. Generically, the electrons lose energy rapidly via synchrotron radiation. These radiated photons provide a target for the accelerated protons, which results in the production of pions, muons and ultimately, neutrinos in the ratio ${\\nu_e:\\nu_\\mu:\\nu_\\tau = 1:2:0}$. The detection and study of ultra-high energy (UHE) neutrinos is thus a probe of the origin of CR and the physics of UHE astrophysical accelerators. As is well understood due to a wealth of data from solar, atmospheric, reactor and accelerator experiments, neutrinos have mass and oscillate into one another. These experiments have led to determinations, to various degrees of accuracy, of the neutrino mass (squared) differences $ \\Delta m^2_{ij} $ between mass eigenstates $ i $ and $ j $, and the mixing angles $ U_{\\alpha i} $ \\cite{Schwetz:2009zz}. The latter characterize the overlap of neutrino mass (or propagation) eigenstates (denoted by $\\nu_i, i=1,2,3$) and the interaction eigenstates (denoted by $\\nu _\\alpha, \\alpha=e,\\mu,\\tau$). Over the very large distances traversed by neutrinos from the most energetic extragalactic hadronic accelerators, the initial source ratios ${\\nu_e^s:\\nu_\\mu^s:\\nu_\\tau^s = 1:2:0}$ will, due to (vacuum) oscillations, transmute, by the time they reach a terrestrial detector, into $\\nu_e^d:\\nu_\\mu^d:\\nu_\\tau^d = 1:1:1$ \\cite{Learned:1994wg,Athar:2000yw}. A series of papers \\cite{Beacom:2002vi,Beacom:2003nh,Barenboim:2003jm,Keranen:2003xd,Beacom:2003eu,Hooper:2004xr,Hooper:2005jp, Meloni:2006gv,Xing:2008fg,Esmaili:2009dz} over the past few years have demonstrated that if the flavour ratios $\\nu_e^d:\\nu_\\mu^d:\\nu_\\tau^d$ detected by extant and upcoming neutrino telescopes were to deviate significantly from this democratic prediction, then important conclusions about physics beyond the Standard Model and neutrino oscillation parameters may consequently be inferred. In addition, deviations of these measured ratios have been shown to be sensitive to neutrino oscillation parameters \\cite{Farzan:2002ct, Beacom:2003zg, Serpico:2005sz, Xing:2006xd, Rodejohann:2006qq, Winter:2006ce, Awasthi:2007az, Lipari:2007su, Blum:2007ie, Pakvasa:2007dc, Choubey:2008di, Esmaili:2009dz} (\\textit{e.g.}~the mixing angles and the Dirac CP violating phase). In this paper we study the spectral distortions in the {\\it diffuse} ({\\it i.e.} integrated over source distribution and redshift) UHE neutrino flux as a probe for the effects of new physics. For specificity, we focus on AGN fluxes, and use, as a convenient bench-mark, the well-known upper bounds first derived by Waxman and Bahcall (WB) \\cite{Waxman:1998yy} and later by Mannheim, Protheroe and Rachen (MPR) \\cite{Mannheim:1998wp} on such fluxes for both neutron-transparent and neutron-opaque sources (or, equivalently, sources that are \\emph{optically thin} and \\emph{optically thick}, respectively, to the emission of neutrons). In particular we focus on the upper bounds to the diffuse neutrino flux from hadronic photoproduction in AGN's derived in \\cite{Mannheim:1998wp} using the experimental upper limit on cosmic ray protons. All distortions in the fluxes are, as would be expected, transmitted to the upper bounds, thus providing a convenient way of representing and studying them. Prior to this, we first demonstrate (in Sec.~\\ref{sec:spectral-averaging}) that the usual (SM) neutrino oscillations not only tend to equilibrate widely differing source flux magnitudes between flavours, but also massage them into a common spectral shape, as one would intuitively expect. Thus observed relative spectral distortions among flavours are a probe of new physics present in the propagation equation. To demonstrate our approach we then focus on two specific examples, \\textit{i.e.,} \\begin{inparaenum}[\\itshape a\\upshape)] \\item Lorentz violation (LV), and \\item Incomplete decay of neutrinos \\end{inparaenum} in the neutrino sector. Our method can straightforwardly be applied to other new physics scenarios and our results translated into bounds on muon track versus shower event rates\\footnote{ These count the sum of \\begin{inparaenum}[\\itshape a\\upshape)] \\item neutral current (NC) events of all flavours, \\item electron neutrino charged current (CC) events at all energies, and \\item $\\nu_\\tau$ induced CC events at energies below $\\leq$ 1 PeV ($10^6$ GeV), \\end{inparaenum} whereas muon track events arise from $\\nu_\\mu$ induced muons born in CC interactions.} for UHE experiments. In our calculations we use the current best-fit values of neutrino mixing paramenters as given in \\cite{Maltoni:2008ka}. The mass squared differences and mixing angles are \\begin{gather*} \\begin{split} \\Delta m_{21}^2 &= 7.65 \\times 10^{-5} \\text{ eV}^2 \\\\ \\Delta m_{31}^2 &= \\pm 2.40 \\times 10^{-3} \\text{ eV}^2 \\end{split} \\\\ \\sin^2(\\theta_{12})=0.321,\\;\\; \\sin^2(\\theta_{23})=0.47,\\;\\; \\sin^2(\\theta_{13})=0.003. \\end{gather*} ", "conclusions": "The detection of UHE neutrinos is imminent. Several detectors will progressively sharpen their capabilities to detect neutrino flavours, beginning with ICECUBE's ability to separate muon tracks from shower events. We have shown that spectral changes in diffuse UHE neutrino fluxes of different flavours are probes of new physics entering the oscillation probability. For specificity we have used AGN sources, and calculated the changes induced in the well-known MPR bounds on both neutron-opaque and neutron-transparent sources. Our calculations can, in a straightforward manner, be repeated for other sources, or represented in terms of the WB bounds or in terms of actual fluxes and event rates. We have shown that diffuse UHE fluxes are sensitive to the presence, over a relatively broad range, of tiny LV terms in the effective oscillation hamiltonian, which signal important breakdowns of pillars of presently known physics. Additionally, in the specific example of neutrino decay, UHE detectors can probe the unexplored range $ 10^{-3} \\text{ s/eV }$ $< \\tau/m < 10^4 \\text{ s/eV } $ via spectral differences in flavours, and thus extend our knowledge of this important parameter. Flavour spectra in decay scenarios also differ depending on the mass hierarchy, and provide a potent tool to probe it. However, it must be kept in mind that the ability to distinguish between flavours in most present-day UHE detectors is as yet not firmly established. At present, detectors are capable of separating between shower events and muon-tracks which include a sum of contributions from each of the flavours rather than dinstinguishing between the individual contributions themselves. Further, the energy resolutions of present or near-future large-volume detectors like the IceCube are not very high. Consequently, deviation of the flavour spectra from the standard predictions has to be significantly large, both in magnitude and shape, for detectability in any of these detectors. Our calculations show that it is indeed possible for such large deviation to occur in the diffuse fluxes of the three flavours for suitable values of the parameters in the case of Lorentz violation as well as decay with inverted hierarchy. Other physics scenarios to which this method may be effectively applied are the existence of pseudo-dirac states, CP violation and quantum decoherence. Although the detection of the effects discussed here will undoubtedly be challenging, given the fundamental nature of the physics to be probed, it would certainly be worthwhile. \\textbf{\\textit{Acknowledgment:}} RG would like to thank Nicole Bell, Maury Goodman, Francis Halzen, Chris Quigg, Georg Raffelt, Subir Sarkar and Alexei Smirnov for helpful discussions, and the theory groups at Fermilab and Brookhaven National Lab for hospitality while this work was in progress." }, "0910/0910.4958_arXiv.txt": { "abstract": "{We present new measurements of the large-scale bulk flows of galaxy clusters based on 5-year WMAP data and a significantly expanded X-ray cluster catalogue. Our method probes the flow via measurements of the kinematic Sunyaev-Zeldovich (SZ) effect produced by the hot gas in moving clusters. It computes the dipole in the cosmic microwave background (CMB) data at cluster pixels, which preserves the SZ component while integrating down other contributions. Our improved catalog of over 1,000 clusters enables us to further investigate possible systematic effects and, thanks to a higher median cluster redshift, allows us to measure the bulk flow to larger scales. We present a corrected error treatment and demonstrate that the more X-ray luminous clusters, while fewer in number, have much larger optical depth, resulting in a higher dipole and thus a more accurate flow measurement. This results in the observed correlation of the dipole derived at the aperture of zero monopole with the monopole measured over the cluster central regions. This correlation is expected if the dipole is produced by the SZ effect and cannot be caused by unidentified systematics (or primary cosmic microwave background anisotropies). We measure that the flow is consistent with approximately constant velocity out to at least $\\simeq$800 Mpc. The significance of the measured signal peaks around 500 $h_{\\rm 70}^{-1}$Mpc, most likely because the contribution from more distant clusters becomes progressively more diluted by the WMAP beam. We can, however, at present not rule out either that these more distant clusters simply contribute less to the overall motion.} ", "introduction": "KABKE1,2 and this work utilize a method of Kashlinsky \\& Atrio-Barandela (2000, hereafter KA-B), which measures CMB dipole at the locations of X-ray clusters. When averaged over many isotropically distributed clusters moving at a significant bulk flow with respect to the CMB, the kinematic term dominates the SZ signal, thereby enabling a measurement of $V_{\\rm bulk}$ over that distance. In Atrio-Barandela et al (2008, hereafter AKKE) we demonstrated that 1) the thermal SZ (TSZ) signal from clusters extends well beyond the measured X-ray extent $\\Theta_{\\rm X}$, and 2) the intra-cluster gas distribution is well approximated by the NFW profile (Navarro et al 1996) expected for dark matter in a $\\Lambda$CDM model. The temperature $T_X$ of hot gas distributed according to these profiles decreases significantly from the cluster cores to the cluster outskirts (Komatsu \\& Seljak 2001), consistent with current measurements (Pratt et al 2007) and numerical simulations (e.g. Borgani et al 2004). Consequently, the monopole produced by the TSZ component ($\\propto \\tau T_X$) decreases, as we increase the cluster aperture, whereas the dipole due to the kinematic SZ (KSZ) component ($\\propto \\tau$, the optical depth due to Thomson scattering) remains measurable out to the aperture where we still detect the TSZ decrement in unfiltered maps (KABKE2). As in KABKE1,2 our dipole coefficients are normalized such that the dipole power $C_{1}$ due to a coherent motion at velocity $V_{\\rm bulk}$ is $C_{1,{\\rm kin}}= T_{\\rm CMB}^2 \\langle \\tau \\rangle^2 V_{\\rm bulk}^2/c^2$, where $T_{\\rm CMB} =2.725$K. We adopt $\\Omega_{\\rm total}=1, \\Omega_\\Lambda=0.7, H_0=70h_{70}$ km/sec/Mpc. To improve upon the all-sky cluster catalogue of Kocevski \\& Ebeling (2006) used by KABKE1,2, we have screened the ROSAT Bright-Source Catalogue (Voges et al.\\ 1999) using the same X-ray selection criteria (including a nominal flux limit of $1\\times 10^ {-12}$ erg/sec, 0.1--2.4 keV) as employed during the Massive Cluster Survey (MACS, Ebeling et al.\\ 2001), as well as the same optical follow-up strategy. Unlike MACS, we apply neither a declination nor a redshift limit though, thereby creating an all-sky list of cluster candidates that extends to three times fainter X-ray fluxes than in KABKE1,2. Optical follow-up observations of clusters identified in this manner and lacking spectroscopic redshifts in the literature (and MACS) are well underway using telescopes on Mauna-Kea/Hawai`i and La-Silla/Chile. As a result, our interim all-sky cluster catalogue currently comprises in excess of 1,400 X-ray selected clusters, {\\it all} of them with spectroscopic redshifts. X-ray properties of all clusters (most importantly total luminosities and central electron densities) are computed as before (KABKE2). Within the same $z$-range ($z\\la0.25$) as our previous study, our new catalog comprises 1,174 clusters outside the KP0 CMB mask. To eliminate low-mass galaxy groups we require that clusters feature $L_X\\geq 2 \\times 10^{43}$ erg/sec; 985 systems meet this criterion. This sample represents a significant improvement over the one in KABKE1,2 largely because of the substantial increase ($z_{\\rm median}<0.1$ in KABKE1,2 vs $z_{\\rm median}\\simeq 0.2$ below) in median cluster redshift and the higher fraction of intrinsically very X-ray luminous systems (Fig. 1a). We applied this new cluster sample to the 5-year WMAP CMB data, processed as described in detail in KABKE1,2. The all-sky dipole in the foreground-cleaned maps from http://www.lambda.gsfc.nasa.gov was removed, and standard CMB masking was applied. We also need to remove the primary CMB fluctuations, produced at last scattering, as they are highly correlated and would contribute significantly to the measured dipole. To this end, all maps were filtered as in KABKE1,2 with a filter that minimizes $\\langle (\\delta T - \\delta T_{\\Lambda{\\rm CDM}})^2\\rangle$. The error budget associated with our filtering is discussed by Atrio-Barandela et al (2010, hereafter AKEKE). Measurement errors were computed as in KABKE1,2 with one important correction. Although the filtering removes much of the CMB fluctuations, a residual component remains, due to cosmic variance and imperfections of the theoretical model. Since this residual is common to all WMAP bands, a component of the errors is correlated between the various DA maps. We address this issue in the following manner. For each of the eight differential assemblies (DAs) we simulate 4,000 realizations with $N_{\\rm cl}=$100--1,000 randomly selected pseudo-clusters outside of the cluster pixels and the CMB mask. In each realization we select the {\\it same} pseudo-clusters for all DAs and evaluate the {\\it mean} monopole and dipole averaged over all DAs: $\\bar{a}_0, \\bar{a}_{1m}$. For the 4,000 realizations at each $N_{\\rm cl}$ we compute the mean and dispersion of $\\bar{a}_0, \\bar{a}_{1m}$ over all the realizations. The distributions have zero mean and their dispersion gives errors which scale as $N_{\\rm cl}^{-1/2}$. We find to good accuracy that the distribution of the simulated dipoles (and monopoles) is Gaussian and the errors on each of the averaged dipole components are $\\sigma_{1m} \\simeq 15 \\sqrt{3/N_{\\rm cl}}\\mu$K and on the monopole $\\sigma_0 \\simeq 15 \\sqrt{1/N_{\\rm cl}} \\mu$K as explained in great detail in AKEKE; the errors of the $x/z$ dipole component are slightly larger/smaller because of the Galactic mask (Fig. 1b). ", "conclusions": "We find a high likelihood of the existence of a coherent bulk flow extending to at least $z\\simeq 0.2$ with an amplitude and in a direction which are in good agreement with our earlier measurements. Our result constitutes a significant improvement in that it extends our previous work to approximately twice the distance accessible to KABKE1,2, supporting their hypothesis that the flow likely extends across much (or all) of the Hubble volume. The flow's axis is also consistent with earlier measurements of the local cluster dipole (Kocevski et al.\\ 2004) as well as with independent measurements of bulk flows on smaller scales by Watkins et al (2009). The velocity reported there is smaller than the numbers in Table 1, although the two amplitudes agree at $< 2$-$\\sigma$ level. Agreement between the two sets of central values would require $\\sqrt{\\langle C_{1,100}\\rangle} \\sim$0.4-0.5$\\mu$K, or a reduction by a factor of $\\sim2$ from unfiltered values for NFW cluster profiles. Feldman et al (2009) extend the Watkins et al analysis and find that the absence of shear in their flow at $\\la$50-100Mpc is consistent with the KABKE1 suggestion of the attractor at superhorizon distances. Fig. \\ref{fig:f2} displays the results obtained in this study compared to expectations from the concordance $\\Lambda$CDM model for 95\\% of cosmic observers. These results cast doubt on the notion that gravitational instability from the observed mass distribution is the sole -- or even dominant -- cause of the detected motion. If the current picture is confirmed, it will have profound implications for our understanding of the global structure of space-time and our Universe's place in it. We acknowledge NASA NNG04G089G/09-ADP09-0050 and FIS2006-05319/GR-234 grants from Spanish Ministerio de Educaci\\'on y Ciencia/Junta de Castilla y Le\\'on. {\\bf REFERENCES}\\\\ Afshordi, N. Geshnizjahi, G. \\& Khoury, J. 2009, JCAP, 8, 30\\\\ Atrio-Barandela, F., Kashlinsky, A., Kocevski, D. \\& Ebeling, H. 2008, Ap.J. (Letters), 675, L57 (AKKE)\\\\ Atrio-Barandela, F., Kashlinsky, A., Ebeling, H., Kocevski, D. \\& Edge, A. 2010, ApJ, submitted, arXiv:1001.1261 (AKEKE)\\\\ Borgani, S. et al 2004, Mon. Not. R. Astron. Soc., 348, 1078\\\\ Carrol, S. et al 2008, arxiv:0811.1086\\\\ Ebeling, H.; Edge, A. C.; Henry, J. P., 2001, ApJ, 553, 668\\\\ Feldman, H., Watkins, R. \\& Hudson, M.J. 2009, arXiv:0911.5516v1\\\\ Gorski, K. et al 2005, Astrophys. J., 622, 759\\\\ Grishchuk, L. P. 1992, Phys. Rev. D 45, 4717\\\\ Kashlinsky, A., Tkachev, I., Frieman, J. 1994, Phys. Rev. Lett., 73, 1582\\\\ Kashlinsky, A. \\& Atrio-Barandela, F. 2000, Astrophys. J., 536, L67 (KA-B)\\\\ Kashlinsky, A., Atrio-Barandela, F., Kocevski, D. \\& Ebeling, H. 2008, Ap.J., 686, L49 (KABKE1)\\\\ Kashlinsky, A., Atrio-Barandela, F., Kocevski, D. \\& Ebeling, H. 2009, Ap.J., 691, 1479 (KABKE2)\\\\ Kashlinsky, A. \\& Jones, B.J.T. 1991, Nature, 349, 753\\\\ Khoury, J. \\& Wyman, M. 2009, PhRevD, 80, 064023\\\\ Kocevski, D.D., Mullis, C.R., \\& Ebeling, H. 2004, Astrophys. J., 608, 721\\\\ Kocevski, D.D. \\& Ebeling, H. 2006, Astrophys. J., 645, 1043\\\\ Komatsu, E. \\& Seljak, U. 2001, Mon. Not. R. Astron. Soc., 327, 1353\\\\ Mersini-Houghton, L. \\& Holman, R. 2009, JCAP, 2, 6\\\\ Navarro, J.F., Frenk, C.S. \\& White, S.D.M. 1996, Astrophys. J., 462, 563\\\\ Pratt, G. et al 2007, Astron. Astrophys. 461, 71\\\\ Turner, M. S. 1991, Phys.Rev., 44, 3737\\\\ Watkins, R., Feldman, H. A. \\& Hudson, M. J. 2009, MNRAS, 392, 743\\\\ Voges, W. et al 1999, A\\&A, 349, 389\\\\ \\clearpage \\thispagestyle{empty} \\begin{deluxetable}{l c c c c c | c | c c | c c c c c} \\tablewidth{0pt} \\tabletypesize{\\scriptsize} \\rotate \\tablecaption{RESULTS \\label{table} } \\tablehead{ \\colhead{(1)} & \\colhead{(2)} & \\colhead{(3)} & \\colhead{(4)} & \\colhead{(5)} & \\colhead{(6)} & \\colhead{(7)} & \\colhead{(8)} & \\colhead{ } & \\colhead{ } & \\colhead{ } & \\colhead{(9)} & \\colhead{ } & \\colhead{ } \\\\ } \\startdata $z\\leq$ & $L_X$-bin & $N_{\\rm cl}$ & $z_{\\rm mean}/z_{\\rm median}$ & $\\bar{a}_{1,x}, \\bar{a}_{1,y}, \\bar{a}_{1,z}$ & $\\sqrt{C_1}$ & $\\langle\\tau_0\\rangle$ & $\\sqrt{C_{1,100}}$ & ($\\mu$K) & & & $\\bar{a}_0$ & ($\\mu$K) & \\\\ & $10^{44}$ erg/s & & & $\\mu$K & $\\mu$K & $\\times10^{-3}$ & 5$^\\prime$ & $\\Theta_{\\rm X}$ & $10^\\prime$ & $15^\\prime$ & $20^\\prime$ & $25^\\prime$ & $30^\\prime$\\\\ \\hline 0.12$^{*}$ & 0.2--0.5 & 142 & 0.061/0.060 & $-4.2\\pm2.7, -0.7\\pm2.3, 0.5\\pm2.3$ & $4.3\\pm2.7$ & 2.8 & 0.2301 & 0.1942 & --2.8 & 0.1 & -- & -- & -- \\\\ 0.12 & 0.5--1 & 194 & 0.081/0.082 & $-2.7\\pm2.3, -2.3\\pm2.0, 1.4\\pm2.0$ & $3.9\\pm2.2$ & 3.5 & 0.2989 & 0.2561 & --2.4 & --1.2 & --0.1 & 0.6 & 0.8 \\\\ 0.12 & $>1$ & 180 & 0.083/0.086 & $4.9 \\pm 2.4, -4.5 \\pm 2.1, 1.5\\pm 2.0$ & $6.8\\pm2.2$ & 5.4 & 0.4610 & 0.3496 & --11.1 & --6.5 & --3.1 & --0.8 & 0.5 \\\\ \\hline \\\\ \\multicolumn{14}{l}{\\vspace{0.1in} $d\\sim 250-370h_{70}^{-1}$Mpc; $(V_x,V_y,V_z)= (174 \\pm 407, -849 \\pm 351 , 348 \\pm 342 )\\times\\frac{+0.3\\mu K}{\\sqrt{\\langle{C}_{1,100}\\rangle}}$ km/sec; $V_{\\rm Bulk}=(934 \\pm 352)\\times\\frac{0.3\\mu K}{\\sqrt{\\langle{C}_{1,100}\\rangle}}$ km/sec; $(l_0,b_0)= (282\\pm 34, 22 \\pm 20)^\\circ$} \\\\ \\hline 0.16 & 0.5--1 & 226 & 0.089/0.087 & $-1.5\\pm2.2, -0.6 \\pm1.9, 2.1 \\pm 1.8$ & $2.7\\pm1.9$ & 3.5 & 0.2843 & 0.2363 & --2.8 & --1.8 & --0.6 & 0.2 & 1.6 \\\\ 0.16 & 1--2 & 191 & 0.106/0.107 & $1.9\\pm2.3, -2.8 \\pm 2.0, -0.5 \\pm 2.0$ & $4.1\\pm2.2$ & 4.4 & 0.3480 & 0.2894 & --4.9 & --1.4 & 0.4 & 1.3 & 1.8 \\\\ 0.16 & $>2$ & 130 & 0.115/0.125 & $4.2\\pm2.8, -8.0\\pm2.4, 4.9\\pm2.4$ & $10.3\\pm2.5$ & 6.8 & 0.4930 & 0.4238 & --11.7 & --7.1 & --2.9 & --0.3 & 0.8 \\\\ \\hline \\\\ \\multicolumn{14}{l}{$^{(a)}$ $d\\sim 370-540 h_{70}^{-1}$Mpc; $(V_x,V_y,V_z) = (410 \\pm 379 , -1,012 \\pm 326, 566 \\pm 319 )\\times\\frac{+0.3\\mu K}{\\sqrt{\\langle{C}_{1,100}\\rangle}}$ km/sec; $V_{\\rm Bulk}=(1,230\\pm 331 )\\times\\frac{0.3\\mu K}{\\sqrt{\\langle{C}_{1,100}\\rangle}}$ km/sec; $(l_0,b_0)=( 292\\pm21, 27 \\pm 15)^\\circ$} \\\\ \\multicolumn{14}{l}{\\vspace{0.1in} $^{(b)}$ $d\\sim 370-540 h_{70}^{-1}$Mpc; $(V_x,V_y,V_z)= (428 \\pm 375 , -1,029 \\pm 323, 575 \\pm 316 )\\times\\frac{+0.3\\mu K}{\\sqrt{\\langle{C}_{1,100}\\rangle}}$ km/sec; $V_{\\rm Bulk}=(1,254\\pm 328 )\\times\\frac{+0.3\\mu K}{\\sqrt{\\langle{C}_{1,100}\\rangle}}$ km/sec; $(l_0,b_0)=(293\\pm 20, 27\\pm 15)^\\circ$} \\\\ \\hline 0.20 & 0.5--1 & 238 & 0.093/0.089 & $-2.5\\pm2.1, -1.3 \\pm 1.8, 1.0 \\pm 1.8$ & $3.0\\pm2.0$ & 3.5 & 0.2828 & 0.2390 & --2.9 & --2.2 & --1.1 & --0.3 & --0.2 \\\\ 0.20 & 1--2 & 248 & 0.122/0.123 & $0.1\\pm2.0, -1.8 \\pm 1.8, -0.3\\pm 1.7$ & $1.8\\pm1.8$ & 4.4 & 0.3231 & 0.2835 & --5.1 & --1.8 & --0.3 & 0.5 & 1.0 \\\\ 0.20 & $>2$ & 208 & 0.140/0.151 & $3.6\\pm2.2, -5.8 \\pm1.9, 4.5\\pm1.9$ & $8.1\\pm2.0$ & 6.6 & 0.4644 & 0.4218 & --9.3 & --5.5 & --1.9 & 0.4 & 1.1 \\\\ \\hline \\\\ \\multicolumn{14}{l}{$^{(a)}$ $d\\sim 380-650 h_{70}^{-1}$Mpc; $(V_x,V_y,V_z)= (213 \\pm 341, -872\\pm 294, 529 \\pm 287)\\times\\frac{+0.3\\mu K}{\\sqrt{\\langle{C}_{1,100}\\rangle}}$ km/sec; $V_{\\rm Bulk}=(1,042 \\pm 295)\\times\\frac{0.3\\mu K}{\\sqrt{\\langle{C}_{1,100}\\rangle}}$ km/sec; $(l_0,b_0)=(284\\pm 24, 30 \\pm 16)^\\circ$} \\\\ \\multicolumn{14}{l}{\\vspace{0.1in} $^{(b)}$ $d\\sim 380-650 h_{70}^{-1}$Mpc; $(V_x,V_y,V_z)= (248 \\pm 337, -880\\pm 291, 538 \\pm 284)\\times\\frac{0.3\\mu K}{\\sqrt{\\langle{C}_{1,100}\\rangle}}$ km/sec; $V_{\\rm Bulk}=(1,061 \\pm 292)\\times\\frac{0.3\\mu K}{\\sqrt{\\langle{C}_{1,100}\\rangle}}$ km/sec; $(l_0,b_0)=(286\\pm 23, 30 \\pm 15)^\\circ$} \\\\ \\hline 0.25 & 0.5--1 & 240 & 0.094/0.090 & $-2.3\\pm2.1, -1.1 \\pm 1.8, 0.9 \\pm 1.8$ & $2.7\\pm2.0$ & 3.5 & 0.2848 & 0.2444 & --2.8 & --2.1 & --1.0 & --0.3 & --0.1 \\\\ 0.25 & 1--2 & 276 & 0.133/0.133 & $-0.2 \\pm 2.0, -1.4\\pm 1.7, 0.7 \\pm 1.6$ & $1.6\\pm1.7$ & 4.4 & 0.3162 & 0.2806 & --5.8 & --2.3 & --0.8 & -0.1 & 0.3 \\\\ 0.25 & $>2$ & 322 & 0.169/0.176 & $3.7\\pm1.8,-4.1\\pm1.5,4.1\\pm1.5$ & $6.9\\pm1.6$ & 6.6 & 0.4434 & 0.4160 & --6.9 & --4.6 & --2.3 & --0.6 & 0.2 \\\\ \\hline \\\\ \\multicolumn{14}{l}{$^{(a)}$ $d\\sim 385-755 h_{70}^{-1}$Mpc; $(V_x,V_y,V_z)= (313\\pm 308, -707\\pm 265, 643\\pm 259)\\times\\frac{+0.3\\mu K}{\\sqrt{{\\langle{C}}_{1,100}\\rangle}}$ km/sec; $V_{\\rm Bulk}=(1,005\\pm267)\\times\\frac{0.3\\mu K}{\\sqrt{\\langle{C}_{1,100}\\rangle}}$ km/sec; $(l_0,b_0)=(296 \\pm 29, 39 \\pm 15)^\\circ$} \\\\ \\multicolumn{14}{l}{\\vspace{0.1in} $^{(b)}$ $d\\sim 385-755 h_{70}^{-1}$Mpc; $(V_x,V_y,V_z)= (352\\pm 304, -713\\pm 262, 652\\pm 256)\\times\\frac{+0.3\\mu K}{\\sqrt{{\\langle{C}}_{1,100}\\rangle}}$ km/sec; $V_{\\rm Bulk}=(1,028\\pm265)\\times\\frac{0.3\\mu K}{\\sqrt{\\langle{C}_{1,100}\\rangle}}$ km/sec; $(l_0,b_0)=(296 \\pm 28, 39 \\pm 14)^\\circ$} \\\\ \\hline \\enddata \\tablecomments{ } \\end{deluxetable} \\clearpage \\begin{figure} \\plotone{f1.eps} \\caption{ \\small{ }} \\label{fig:f1} \\end{figure} \\clearpage \\begin{figure} \\plotone{f2.eps} \\caption{ \\small{ }} \\label{fig:f2} \\end{figure} \\clearpage" }, "0910/0910.1099_arXiv.txt": { "abstract": "We present observations at 1.2~mm with MAMBO-II of a sample of $z\\grtsim2$ radio-intermediate obscured quasars, as well as CO observations of two sources with the Plateau de Bure Interferometer. The typical rms noise achieved by the MAMBO observations is 0.55~mJy beam$^{-1}$ and 5 out of 21 sources (24\\%) are detected at a significance of $\\geq3\\sigma$. Stacking all sources leads to a statistical detection of $\\langle S_{1.2~\\rm mm} \\rangle = 0.96\\pm0.11$~mJy and stacking only the non-detections also yields a statistical detection, with $\\langle S_{1.2~\\rm mm} \\rangle = 0.51\\pm0.13$~mJy. At the typical redshift of the sample, $z=2$, 1~mJy corresponds to a far-infrared luminosity \\lfir$\\sim4\\times10^{12}$~\\lsol. If the far-infrared luminosity is powered entirely by star-formation, and not by AGN-heated dust, then the characteristic inferred star-formation rate is $\\sim$700~\\msolyr. This far-infrared luminosity implies a dust mass of \\mdust$\\sim3\\times10^{8}$~\\msol, which is expected to be distributed on $\\sim$kpc scales. We estimate that such large dust masses on kpc scales can plausibly cause the obscuration of the quasars. Combining our observations at 1.2~mm with mid- and far-infrared data, and additional observations for two objects at 350~\\mum\\, using SHARC-II, we present dust SEDs for our sample and derive a mean SED for our sample. This mean SED is not well fitted by clumpy torus models, unless additional extinction and far-infrared re-emission due to cool dust are included. This additional extinction can be consistently achieved by the mass of cool dust responsible for the far-infrared emission, provided the bulk of the dust is within a radius $\\sim$2-3~kpc. Comparison of our sample to other samples of $z\\sim2$ quasars suggests that obscured quasars have, on average, higher far-infrared luminosities than unobscured quasars. There is a hint that the host galaxies of obscured quasars must have higher cool-dust masses and are therefore often found at an earlier evolutionary phase than those of unobscured quasars. For one source at $z=2.767$, we detect the CO(3-2) transition, with $S_{\\rm CO}\\Delta \\nu=$630$\\pm$50 mJy~\\kms, corresponding to $L_{\\rm CO (3-2)}=$3.2$\\times10^{7}$~\\lsol, or a brightness-temperature luminosity of $L'_{\\rm CO (3-2)}=2.4\\times10^{10}$ K~\\kms~pc$^{2}$. For another source at $z=4.17$, the lack of detection of the CO(4-3) line suggests the line to have a brightness-temperature luminosity $L'_{\\rm CO (4-3)}<1\\times10^{10}$ K~\\kms~pc$^{2}$. Under the assumption that in these objects the high-J transitions are thermalised, we can estimate the molecular gas contents to be $M_{\\rm H_{2}}=1.9\\times10^{10}$ \\msol and $<8\\times10^{9}$ \\msol, respectively. The estimated gas depletion timescales are $\\tau_{\\rm g}=4$~Myr and $<$16~Myr, and low gas-to-dust mass ratios of \\mgas$/$\\mdust$=19$ and $\\leq8$ are inferred. These values are at the low end but consistent with those of other high-redshift galaxies. ", "introduction": "Quasars are believed to be powered by supermassive black holes (SMBHs) rapidly growing by accreting matter at a large fraction of their Eddington rate\\footnote{In this article we use the term quasar for the most powerful active galactic nuclei (AGNs). In the case of unobscured AGNs, the dividing line is taken at a B-band magnitude of $M_{\\rm B}=-23.5$ which, assuming the typical quasar SED of \\citet{1994ApJS...95....1E}, corresponds to $L_{\\rm bol}\\grtsim 2\\times10^{12}$~\\lsol. We will therefore also refer to obscured AGNs as obscured quasars if they have $L_{\\rm bol}\\grtsim 2\\times10^{12}$~\\lsol.}. The accreted matter forms a disk, is heated to temperatures of $\\sim10^{3}-10^{5}$~K, and emits thermal emission at optical and ultraviolet wavelengths. Clouds of gas ionised by this radiation will emit line radiation, and lines which are relatively close to the black hole have rapid orbits which will show broad velocity dispersions ($>$2000 \\kms), while clouds further away will show narrower lines. Dust can survive when the equilibrium temperature with the ultraviolet photons is $\\lesssim$2000~K. This dust absorbs optical and ultraviolet radiation and re-emits it at infrared wavelengths characteristic of the dust temperature. If the line of sight of an observer to the central region is blocked by dust, the optical and ultraviolet continuum, as well as the broad lines, will not be observable. Such a source is known as an obscured quasar. While radio-loud obscured quasars have been known for a long time, in the form of high-excitation narrow-line radio galaxies \\citep[see ][ for a review]{1995PASP..107..803U}, the population of radio-quiet obscured quasars had remained elusive, with only a few individual objects known. This has changed recently, when large numbers of these sources have been found in the spectroscopic database of the Sloan Digital Sky Survey \\citep{2003AJ....126.2125Z}. Samples of spectroscopically-confirmed obscured quasars have now also been identified in surveys using X-ray selection \\citep[e.g.][]{2004ApJS..155..271S,2005AJ....129..578B} as well as mid-infrared, sometimes combined with radio, selection \\citep[e.g.][]{2005ApJ...622L.105H,2005Natur.436..666M,2007ApJ...669L..61L}. It seems that the obscured quasars are at least as common as the the unobscured quasars \\citep[e.g.][]{2008AJ....136.2373R}, and probably outnumber these by a $\\sim$2:1 ratio \\citep[e.g.][]{2005Natur.436..666M,2008ApJ...674..676M,2007ApJ...669L..61L,2008ApJ...675..960P}. There is a large body of evidence supporting obscuration of quasars as an orientation effect, where obscured and unobscured quasars are intrinsically identical sources, only dust around the accretion disk covers certain lines of sight so that the broad emission lines and the optical and ultraviolet continuum are not visible \\citep[the dusty torus of the unified schemes, e.g.][]{1995PASP..107..803U}. However, dust distributed within the host galaxy can also cause obscuration, particularly if the galaxy is at an early evolutionary phase and is rich in gas and dust \\citep[see e.g.][]{1982ApJ...256..410L,1988ApJ...325...74S,1999MNRAS.308L..39F}. The importance of galaxy-scale dust is evident in recent studies of quasars where obscuration cannot be solely assigned to dust from a torus around the accretion disk: in some objects even the narrow-lines are obscured \\citep{2005Natur.436..666M,2006MNRAS.370.1479M,2006ApJ...645..115R}. The jets emanating from some obscured quasars are face-on \\citep{2006MNRAS.373L..80M,2007ApJ...667L..17S,2009arXiv0905.1605K}, something that is not expected for quasars obscured by the torus, but can be explained if the obscuring dust is on a larger ($\\sim$kpc) scale. The deep silicate absorption features and spectral energy distributions (SEDs) suggest foreground extinction is sometimes present \\citep{2007ApJ...654L..45L,2008ApJ...675..960P}. This all provides indirect evidence suggesting obscuring dust on $\\sim$~kpc scales, something which has been spectroscopically confirmed through the measurement of the H$\\alpha$ to H$\\beta$ Balmer decrement \\citep {2007ApJ...663..204B}. Dust present on kpc scales is expected to be relatively cool ($T\\sim$50~K) and will emit thermally in the far-infrared ($\\lambda \\grtsim$40~\\mum), regardless of whether it is heated by young stars or by the central active galactic nucleus \\citep[AGN, see e.g.][ and Section~\\ref{sec:scale}]{1987ApJ...320..537B,1989ApJ...347...29S,2004A&A...421..129S}. Observing this dust around the peak of its emission (around rest-frame $\\sim$65~\\mum) is not optimal, since warmer dust can still contribute significantly there. The Rayleigh-Jeans tail of the emission, however, will have a much smaller contribution from warm dust, and has the additional advantage of being observable from the ground, for example in the atmospheric windows at 850~\\mum\\, or 1.2~mm. In this paper we present continuum observations of a sample of 21 high-redshift ($z\\grtsim$2) radio-intermediate obscured quasars at 1.2~mm, combined with other infrared and submillimetre data, as well as a search for CO in two of the sources. The sample was selected to approximately match the break in the unobscured luminosity function at $z\\sim2$ \\citep[$M_{\\rm B}\\sim$-25.7, ][]{2004MNRAS.349.1397C}, so in terms of radiation these quasars represent the energetically-dominant population around the peak of the quasar activity. The sources are, however, radio-intermediate in that their radio luminosities are slightly higher than those expected from radio-quiet quasars \\citep[with \\lrada\\, $\\sim10^{24}$~\\whzsr][]{2006MNRAS.373L..80M}. Nevertheless, observations with very large baseline interferometry suggest the relative importance of the compact cores makes them more similar to the genuinely radio-quiet population than to the radio galaxies and radio-loud quasars \\citep[see][]{2009arXiv0905.1605K}. Section~\\ref{sec:obs} summarises the observations and data reduction. In Section~\\ref{sec:res} we show the resulting fluxes, inferred luminosities, star-formation rates, dust masses, characteristic scale of the cool dust and discuss the possible obscuration by kpc-scale dust. Section~\\ref{sec:broad} shows the results of broad-band SED fitting and Section~\\ref{sec:mean} presents the mean SED for our sample. In Section~\\ref{sec:co} we discuss the molecular gas observations. We compare our sample to other millimetre or submillimetre observations of $z\\sim2$ quasars in Section~\\ref{sec:comp}. The discussion and summary are found in Section~\\ref{sec:disc}. ", "conclusions": "\\label{sec:disc} We have observed a sample of $z\\grtsim2$ radio-intermediate obscured quasars at 1.2~mm (250~GHz) using MAMBO. The detection rate is 5 out of 21 (24\\%). Sources with narrow lines seem to have a higher detection rate than sources with no narrow lines, but with such small numbers the difference is not significant. Larger samples will be required to study any differences. Stacking leads to a statistical detection of $\\langle S_{1.2~\\rm mm} \\rangle = 0.96\\pm0.11$. Even if only the non-detections are stacked, they still yield a statistical detection, with $\\langle S_{1.2~\\rm mm} \\rangle = 0.51\\pm0.13$. Thus, the typical flux density of this sample is $\\sim$0.5-1.0~mJy, and this corresponds to a far-infrared luminosity $\\sim4\\times10^{12}$~\\lsol. If the far-infrared luminosity is powered entirely by star-formation, and not by AGN-heated dust, then the typical inferred star-formation rate is $\\sim$700~\\msolyr. This large star-formation rate is comparable to those inferred in ULIRGs and SMGs, the most powerful starburst galaxies known. The observations at 1.2~mm also allow us to estimate the mass of the cool dust, and we find a typical mass of $\\sim3\\times10^{8}$~\\msol. If heated by young stars, this dust is expected to be distributed on $\\sim$2~kpc scales (the scale found for SMGs). If it is heated by the central AGN, with \\lbol$\\sim10^{13}$~\\lsol, then it is expected to be distributed on scales $\\lesssim$10~kpc. Indeed, we estimate that such a large mass of cool dust is capable of causing alone the large values of \\av\\, in this sample, without help of an obscuring torus. Combining our observations at mid-infrared and millimetre wavelengths, we present dust SEDs for our sample, and derive a typical SED for our sample of high-redshift obscured quasars. The clumpy tori models cannot reproduce the SED. However, clumpy tori models with an additional screen of cold dust (and the corresponding far-infrared emission) do reproduce this mean SED. The amount of extinction from this additional screen is derived from the cool dust mass, from the observations at 1.2~mm. The depth of the silicate feature can be consistently achieved by the inferred cool dust mass provided that the dust is on scales $\\lesssim$2~kpc, which is in excellent agreement with the typical radius of the far-infrared emission in SMGs. Again, this lends support to kpc-scale dust along the host galaxy playing an important role in the obscuration of these sources. Obscuration by dust on kpc scales would also explain why in about half of the sample the narrow emission lines from the central AGN are not detectable, as well as why in some cases the jets seem to be pointing towards us, something unexpected if they are only obscured by the torus of the unified schemes. However, we remind the reader that unobscured quasars have also been found to have similar cool-dust masses, and that in our sample there is not a one-to-one correspondence between the presence (or lack of) narrow lines and detection at 1.2~mm. The presence of a large mass of dust does not guarantee obscuration. A clumpy dusty interstellar medium where different lines of sight have vastly different optical depths is a likely explanation for why some quasars with large dust masses are obscured while others are not. When comparing to other samples of high-redshift quasars, we find that the obscured quasars probably have a higher fraction of their luminosity emerging at far-infrared wavelengths, compared to unobscured quasars. The far-infrared emission is isotropic, so that this difference cannot be ascribed to orientation-dependent obscuration: obscured quasars are likely to have higher far-infrared luminosities and cool-dust masses than unobscured quasars, suggesting the host galaxies are in a dustier, presumably earlier, phase. Additionally, we have searched for molecular gas in two sources at $z=2.767$ and 4.169, using the CO (3-2) and (4-3) transitions. In the $z=2.767$ source we detect a line with $L_{\\rm CO (3-2)}=$3.2$\\times10^{7}$~\\lsol\\, (equivalent to a brightness-temperature luminosity of $L'_{\\rm CO (3-2)}=2.4\\times10^{10}$ K~\\kms~pc$^{2}$). In the other source, the lack of detection suggests a line luminosity $L_{\\rm CO (4-3)}<$3$\\times10^{7}$~\\lsol\\, ($L'_{\\rm CO (4-3)}<1\\times10^{10}$ K~\\kms~pc$^{2})$. Under the assumption that in these objects the (3-2) and (4-3) transitions are thermalised, we can estimate the molecular gas contents to be $M_{\\rm H_{2}}=1.9\\times10^{10}$ and $<8\\times10^{9}$ \\msol, respectively. The estimated gas depletion timescales are $\\tau_{\\rm g}=4$ and $<$16~Myr, and low gas-to-dust mass ratios of \\mgas$/$\\mdust$=19$ and $\\leq20$ are inferred. A dynamical mass of $M_{\\rm dyn} {\\rm sin^{2}}i=6\\times10^{9}$~\\msol\\, is estimated from the CO(3-2) detection." }, "0910/0910.4733_arXiv.txt": { "abstract": "{ The solar abundances have undergone a major downward revision in the last decade, reputedly as a result of employing 3D hydrodynamical simulations to model the inhomogeneous structure of the solar photosphere. The very low oxygen abundance advocated by \\citet{asplund04}, A(O)=8.66, together with the downward revision of the carbon and nitrogen abundances, has created serious problems for solar models to explain the helioseismic measurements. In an effort to contribute to the dispute we have re-derived photospheric abundances of several elements independently of previous analysis. We applied a state-of-the art 3D (CO5BOLD) hydrodynamical simulation of the solar granulation as well as different 1D model atmospheres for the line by line spectroscopic abundance determinations. The analysis is based on both standard disc-centre and disc-integrated spectral atlases; for oxygen we acquired in addition spectra at different heliocentric angles. The derived abundances are the result of equivalent width and/or line profile fitting of the available atomic lines. We discuss the different granulation effects on solar abundances and compare our results with previous investigations. According to our investigations hydrodynamical models are important in the solar abundance determination, but are not responsible for the recent downward revision in the literature of the solar metallicity. ", "introduction": "In this work we would like to face the most common questions we are confronted with in the analysis of the photospheric solar abundances: \\begin{itemize} \\item ``Are 3D models important in the abundances determination?'' \\item ``Are 3D models responsible for the downward revision of the solar metallicity?'' \\end{itemize} Our answer to the first question is yes. As we know from previous investigations \\citep{zolfito}, 3D solar metallicity models do not experience the over-cooling in the external layers, not detected in 1D models, that metal-poor 3D models show. One could then expect that 3D models are not fundamental for the solar abundance determinations. If for some elements (P, Eu, Hf) the granulation effects are in fact negligible, this is not the case for others, such as Fe, Th, and also oxygen. On top of that one should not forget that 1D models require some input parameters (mixing-length parameter, \\mlp, and microturbulence, $\\xi_{\\rm micro}$) implicit in 3D models; so that in the case such parameters are fundamental (\\mlp\\ for C, O, and Fe, $\\xi_{\\rm micro}$ for N) 3D models should be highly preferred. \\begin{table} \\caption{Influence of the microturbulence on the 3D corrections.} \\label{microt} { \\begin{tabular}{ccc} \\hline \\noalign{\\smallskip} $\\xi _{\\rm \\xx}$ & \\multicolumn{2}{c}{$\\left({\\rm A(Y)_{\\rm 3D}}-{\\rm A(Y)_{\\rm \\xx}}\\right)$ [dex]} \\\\ \\kms\\ & Flux & Intensity \\\\ \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} 0.6 & +0.036 & - -\\\\ 0.9 & - - & +0.037\\\\ 1.2 & +0.130 & - - \\\\ 1.5 & - - & +0.139\\\\ \\noalign{\\smallskip} \\hline \\end{tabular} } \\end{table} \\begin{table*} \\caption{Photospheric solar abundances.} \\label{sunabbo} \\begin{center} {% \\begin{tabular}{lrlllll} \\hline \\noalign{\\smallskip} EL & N & \\cobold\\ & AG89 & GS98 & AGS05 & AGSS09\\\\ \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} Li & 1 & $1.03\\pm 0.03$ & $1.16\\pm 0.10$ & $1.10\\pm 0.10$ & $1.05\\pm 0.10$ & $1.05\\pm 0.10$ \\\\ C & 43 & $8.50\\pm 0.06$ & $8.56\\pm 0.04$ & $8.52\\pm 0.06$ & $8.39\\pm 0.05$ & $8.43\\pm 0.05$ \\\\ N & 12 & $7.86\\pm 0.12$ & $8.05\\pm 0.04$ & $7.92\\pm 0.06$ & $7.78\\pm 0.06$ & $7.83\\pm 0.05$ \\\\ O & 10 & $8.76\\pm 0.07$ & $8.93\\pm 0.035$& $8.83\\pm 0.06$ & $8.66\\pm 0.05$ & $8.69\\pm 0.05$ \\\\ P & 5 & $5.46\\pm 0.04$ & $5.45\\pm 0.04$ & $5.45\\pm 0.04$ & $5.36\\pm 0.04$ & $5.41\\pm 0.03$ \\\\ S & 9 & $7.16\\pm 0.05$ & $7.21\\pm 0.06$ & $7.33\\pm 0.11$ & $7.14\\pm 0.05$ & $7.12\\pm 0.03$ \\\\ K & 6 & $5.11\\pm 0.09$ & $5.12\\pm 0.13$ & $5.12\\pm 0.13$ & $5.08\\pm 0.07$ & $5.03\\pm 0.09$ \\\\ Fe & 15 & $7.52\\pm 0.06$ & $7.67\\pm 0.03$ & $7.50\\pm 0.05$ & $7.45\\pm 0.05$ & $7.50\\pm 0.04$ \\\\ Eu & 5 & $0.52\\pm 0.03$ & $0.51\\pm 0.08$ & $0.51\\pm 0.08$ & $0.52\\pm 0.06$ & $0.52\\pm 0.04$ \\\\ Hf & 4 & $0.87\\pm 0.04$ & $0.88\\pm 0.08$ & $0.88\\pm 0.08$ & $0.88\\pm 0.08$ & $0.85\\pm 0.04$ \\\\ Os & 3 & $1.36\\pm 0.19$ & $1.45\\pm 0.10$ & $1.45\\pm 0.10$ & $1.45\\pm 0.10$ & $1.40\\pm 0.08$ \\\\ Th & 1 & $0.08\\pm 0.03$ & $0.12\\pm 0.06$ & ${\\it 0.09\\pm 0.02}$ & ${\\it 0.06\\pm 0.05}$ & $0.02\\pm 0.10$ \\\\ & & & & & & \\\\ Z & & 0.0153 & 0.0189 & 0.0171 & 0.0122 & 0.0134 \\\\ Z/X& & 0.0209 & 0.0267 & 0.0234 & 0.0165 & 0.0183 \\\\ \\noalign{\\smallskip} \\hline \\end{tabular} } \\end{center} Note: AG89=\\citet{anders89}, GS98=\\citet{grevesse98}, AGS05=\\citet{sunabboasp}, and AGSS09=\\citet{asplund09}. \\end{table*} Our answer to the second question is no. We would like to remark that this answer depends on the 1D reference model one is considering. For all the elements, except N and Th, our 3D model gives an abundance larger than the 1D model. Th is an exception because its lower 3D abundance is related to the fact that the line is on the red wing of a stronger Fe-Ni blend, and the asymmetry of this blend, correctly taken into account in 3D analysis, lowers the Th contribution \\citep{thhf}. For N all the lines considered are of high excitation energy, so their 1D abundance is very sensitive on the choice of \\mlp. To try to catch the effects of granulation on the abundances determination, we selected 1D models sharing the micro-physics with our 3D solar model atmosphere. But the comparison of abundances derived from 1D and 3D models is not unambiguously defined, relying on the choice of \\mlp\\ and $\\xi_{\\rm micro}$. ", "conclusions": "In the light of our work we think that the use of 3D models in the solar abundance determination is useful. This is for the following reasons: \\begin{itemize} \\item no constraint is necessary on \\mlp\\ and $\\xi_{\\rm micro}$; \\item a difference of few hundreds of dex in the abundance, negligible for the majority of the stellar analysis, is important in the solar context. \\end{itemize} \\balance" }, "0910/0910.1320_arXiv.txt": { "abstract": "We report a relation between radio emission in the inner jet of the Seyfert galaxy \\objectname{3C~120} and optical continuum emission in this galaxy. Combining the optical variability data with multi-epoch high-resolution very long baseline interferometry observations reveals that an optical flare rises when a superluminal component emerges into the jet and its maxima is related to the passage of such component through the location a stationary feature at a distance of $\\approx$1.3\\,parsecs from the jet origin. This indicates that a significant fraction of the optical continuum produced in \\objectname{3C~120} is non-thermal and it can ionize material in a sub-relativistic wind or outflow. We discuss implications of this finding for the ionization and structure of the broad emission line region, as well as for the use of broad emission lines for determining black hole masses in radio-loud AGN. ", "introduction": "In the current astrophysical paradigm for active galactic nuclei (AGN), each constituent of an AGN contributes to a specific domain in the spectral energy distribution (SED). The SED of some AGN, show a significant amount of energy excess at UV/optical wavelengths, which is commonly attributed to be produced in an accretion disk around the putative central supermassive black hole (BH). Surrounding the accretion disk at $\\lesssim 1$ pc, there is a central broad line region (BLR) formed by high density gaseous clouds orbiting the central BH. The thermal UV/optical emission radiated from the disk is thought to be the prime source of variable optical continuum and a dominant factor for the ionization of the BLR material. In radio-loud AGN however, the broad-band continuum can also have a substantial contribution from non-thermal synchrotron radiation generated in the relativistic jet (D'Arcangelo et al. 2007, Marscher et al. 2008, Soldi et al. 2008). To understand the mechanism and properties of the broad-line generation in such objects, it is pivotal to be able to localize and identify the region where the bulk of the non-thermal optical continuum emission is produced. An efficient way to identify a region responsible for the production of the variable optical continuum emission in radio-loud AGN is to combine long-term optical monitoring and regular very long baseline interferometry (VLBI) observations. A link between the optical emission and relativistic jet in 3C~120 was suggested by Belokon (1987), based on a correlation found between the optical outburst and ejections of plasma clouds (jet components) in the jet. Using the radio and optical data available from 14 years monitoring of the radio galaxy 3C\\,390.3, Arshakian et al. (2008, 2009) found that ejections of new components can be associated with optical flares occurring on time scales from months to years and reaching their maxima during passages of the jet components through a stationary emitting region in the inner jet, at a distance of $\\sim 0.5$\\,pc from the jet origin. This indicates that the variable optical continuum emission can be generated in the innermost part of the jet, in the region located upstream from the stationary feature. This has an important implication for the existence of a non-virialized outflowing broad-line region associated with the jet. \\begin{figure*}[t!] \\includegraphics[width=\\textwidth]{f1.eps} \\caption{ Compact jet in 3C~120 observed on January 7, 2005, with the VLBA at 15~GHz (2cm). Shaded ellipse in the lower left corner marks the restoring point-spread function (beam) with the FWHM of 0.53 $\\times$ 1.26 mas oriented at an angle of three degrees (clockwise rotation). The peak flux density in the image is 974 mJy/beam and the rms noise is 0.26 mJy/beam. Contours are drawn at 1, $\\sqrt{2}$, 2...\\% of the lowest contour shown at 0.9 mJy/beam. The jet structure is parametrized by a set of two-dimensional, circular Gaussian features obtained from fitting the visibility amplitudes and phases (Pearson 1997). The model-fit components (C4-C12) located within the inner 14 mas from the core are shown in Figure \\ref{Fig:3c120_kinematic_model}. Label marks show the location of the model-fit components and the two stationary components D and S1 are highlighted by open circles.} \\label{Fig:3c120_jet} \\end{figure*} It should be noted that stationary and low-pattern speed features have been recently reported in parsec-scale jets of a number of prominent radio sources ({\\em e.g.}, Kellermann et al. 2004, Savolainen et al. 2006, Lister et al. 2009b), alongside faster, superluminally moving components embedded in the same flows (Kellermann et al. 2004, Jorstad et al. 2005). Although specific geometric conditions and extremely small viewing angles may lead to formation of such features in relativistic flows (Alberdi et al. 2000), it is more likely that these features represent standing shocks (for instance, recollimation shocks in an initially over-pressurized outflow; G\\'omez et al. 1995, Perucho \\& Mart\\'i 2007). Such standing shocks may play a major role in accelerating particles near the base of the jet (Mandal \\& Chakrabarti 2008; Becker et al. 2008), and could be responsible for the persistent high levels of polarization in blazars (D'Arcangelo et al. 2007; Marscher et al. 2008). It is important to understand whether the correlations found in Arshakian et al. (2008, 2009) are characteristic only for the 3C\\,390.3 or common for all radio galaxies and quasars. In this letter, we combined archival VLBI monitoring data and data available from photometric monitoring of 3C\\,120 to study the link between properties of the superluminal jet and optical continuum flares. Throughout the paper, a flat $\\Lambda$CDM cosmology is assumed, with the Hubble constant H$_{0}= 70$ km s$^{-1}$ Mpc$^{-1}$ and matter density $\\Omega_{m}=0.3$. This corresponds to a linear scale of 0.658\\,pc/mas, at the redshift $z=0.033$ of \\objectname{3C~120}. ", "conclusions": "To investigate the link between optical continuum variability and subparsec-scale jet in the radio-loud galaxy 3C\\,120, we combined the radio VLBI (15\\,GHz) and optical photometry (B-band) data observed during the period of nearly eight years. We found a significant correlation between the jet kinematics on parsec-scales and optical flares on scales from several months to about two years. All optical flares are associated with the jet ejection events: the flare rises after the epoch of ejection of a new jet component (at D) and it reaches the maximum around the epoch at which the ejected radio knot passes the stationary radio component (S1) downstream the jet. The radio passages of new components through S1 delay the maxima of optical flares on the average by $\\la 0.1$ yr. These results confirm correlations found in Arshakian~et~al.~(2008,~2009) for the radio galaxy 3C\\,390.3 and support the idea that the correlation between optical flares and kinematics of the jet could be a common feature for all radio-loud galaxies and quasars. The link between optical continuum variability and kinematics of the parsec-scale jet is interpreted in terms of optical flares generated by disturbances in the jet flow while moving from the stationary component D of the jet to the stationary component S1 located at a distance of about one parsec. Modeling of optical flares as synchrotron flares caused by density variations in the relativistic flow showed that all optical flares require at most a factor of seven increase of the particle density. The variation of the particle density at subparsec-scales should be smaller if consider a more realistic model in which the magnetic field and/or Lorentz factor of the jet increase downstream of the jet, in the acceleration/collimation zone, between D and S1. In this scenario, the correlated X-ray and optical continuum emission are produced in the relativistic jet at some distance ($< 1$ pc) from the central black hole rather than in the accretion disk. \\\\ Establishing such a relation may have strong implications for the physics of the central regions in radio-loud AGN and AGN in general (see also Arshakian et al. 2009). The link between the optical continuum and radio jet challenges the existing models in which the optical continuum and broad-line emission are both localized around the disk or near the central black hole of an AGN. It also questions the common assumptions about the central engine used, in particular, for the reverberation mapping technique (Peterson et al. 2002), which combines the broad emission line width with the size of the line-emitting region estimated from the optical continuum luminosity. In 3C\\,120, a substantial fraction of the ionizing continuum is produced in the relativistic jet and it illuminates thermal material most likely organized in a sub-relativistic outflow. The time delays and profile widths measured in radio-loud AGN may reflect not only the Keplerian motion of the emitting line gas, but also an outflowing component of the gas accelerated by non-gravitational forces. This can lead to large errors in estimates of black hole masses made from monitoring of the broad emission lines. In order to address these issues, we are carrying out a long-term spectropolarimetry campaign for the \\objectname{3C~120} and several other nearby radio galaxies aimed to determine the amount of non-thermal optical continuum and to tackle evidence for non-virial motions in the BLR. Follow-up high resolution radio observations at other frequencies and epochs, can help to characterize the nature of the components D and S1. We thank anonymous referees for a number of useful comments and suggestions which significantly improved this manuscript. This work was supported by CONACYT research grant 54480 (Mexico) and the Russian Foundation for Basic Research (Grant No. 09-02-01136a) . JLT acknowledges support from the CONACYT program for PhD studies, the International Max-Planck Research School for Radio and Infrared Astronomy at the Universities of Bonn and Cologne and the Deutscher Akademischer Austausch Dienst (DAAD) for a short-term scholarship in Germany. This research has made use of data from the MOJAVE database that is maintained by the MOJAVE team (Lister et al. 2009a). The VLBA is an instrument of the National Radio Astronomy Observatory, a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc." }, "0910/0910.1927_arXiv.txt": { "abstract": "Although they are only minor constituents of the interstellar medium, halogen-containing molecules are of special interest because of their unique thermochemistry. Here, we present a theoretical study of the chemistry of interstellar molecules containing the halogen elements chlorine and fluorine. We have modeled both diffuse and dense molecular clouds, making use of updated estimates for the rates of several key chemical processes. We present predictions for the abundances of the three halogen molecules that have been detected to date in the interstellar medium: HF, CF$^+$ and HCl. As in our previous study of fluorine-bearing interstellar molecules, we predict HF to be the dominant gas-phase reservoir of fluorine within both diffuse and dense molecular clouds; we expect the {\\it Herschel Space Observatory} to detect widespread absorption in the HF $J=1-0$ transition. Our updated model now overpredicts the CF$^+$ abundance by a factor $\\simgt 10$ relative to observations of the Orion Bar; this discrepancy has widened because we now adopt a laboratory measurement of the CF$^+$ dissociative recombination rate that is smaller than the estimate we adopted previously. This disagreement suggests that the reaction of C$^+$ with HF proceeds more slowly than the capture rate assumed in our model; a laboratory measurement of this reaction rate would be very desirable. Our model predicts diffuse cloud HCl abundances that are similar to those predicted previously and detected tentatively toward $\\zeta$~Oph. Two additional species are potentially detectable from photodissociation regions: the $\\rm H_2Cl^+$ and $\\rm HCl^+$ molecular ions. Ortho-$\\rm H_2Cl^+$ has its lowest-lying transition in the millimeter spectral region observable from the ground, and the lowest rotational transition of HCl$^+$ is observable with {\\it Herschel}'s HIFI instrument. ", "introduction": "Of the $\\sim$ 150 distinct molecules detected thus far in interstellar clouds and circumstellar outflows, over 90$\\%$ contain only hydrogen and/or elements in groups IVA, VA, and VIA\\footnote{Here we use the traditional designations used in the United States. The current IUPAC designations for groups IVA, VA, and VIA are respectively 14, 15, and 16; those for groups IA, IIA, IIIA and VIIA are 1, 2, 13, and 17} of the periodic table: viz.\\ carbon, nitrogen, oxygen, silicon, phosphorus, and sulfur. Of the remaining 12 detected molecules, eight contain elements in groups IA, IIA and IIIA (viz.\\ sodium, potassium, magnesium and aluminum) and are found exclusively in circumstellar envelopes (e.g.\\ Ziurys 1996, and references therein) with abundances that are set in cool stellar photospheres; one contains iron (FeO, detected toward the Sgr B2 interstellar gas cloud by Walmsley et al.\\ 2002); and the remaining 3 are interstellar molecules containing elements of the halogen group (VIIA): HCl, HF, and CF$^+$. Although they are only minor constituents of the interstellar medium, halogen-containing molecules are of special interest because of their unique thermochemistry. The basic behavior of any atom in the interstellar medium is determined by two or three key thermochemical considerations. First, the first ionization potential (IP) determines whether the atom will be neutral or ionized within diffuse molecular clouds: an atom X with IP(X) $>$ IP(H) = 13.6 eV will be shielded by atomic hydrogen from the interstellar radiation field (ISRF) and will be predominantly neutral, while an atom with IP(X) $<$ IP(H) will be predominantly ionized. Second, the dissociation energy of the neutral hydride, $D_0$(HX), determines whether X can react exothermically with H$_2$: atoms for which $D_0$(HX) $>$ $D_0$(H$_2$) = 4.48 eV can react exothermically with H$_2$ to form HX, while atoms with $D_0$(HX) $ < D_0$(H$_2$) cannot. Third (and of importance mainly for atoms with IP $<$ 13.6 eV), the dissociation energy of the hydride molecular ion, $D_0$(HX$^+$), similarly determines whether the ion X$^+$ can react exothermically with H$_2$ to form HX$^+$. Fluorine and chlorine\\footnote{In this paper we combine our attention to Cl and F; bromine and iodine have much smaller cosmic abundances than chlorine and fluorine, and molecules containing bromine and iodine are unlikely to be detectable with current technology.} are unique in regard to the second and third criteria, each in a different way: F is the only atom -- and Cl$^+$ is the only ion with an appearance potential $<$ 13.6 eV -- that can react exothermically with H$_2$ to form a diatomic hydride. The uniqueness of these elements are presented graphically in Figure 1, which shows $D_0$(HX) for each atom X in the first three rows of the periodic table. The molecular binding energy of HX is largest for fluorine and decreases, nearly monotonically, as one moves leftwards and downwards in the periodic table. In Figure 2, $D_0$(HX) and $D_0$(HX$^+$) are shown for all the elements of which interstellar molecules have been detected. Red squares apply to atoms with IP $<$ IP(H) and blue squares to those with IP $>$ IP(H). With the exception of the halogen elements, all the elements fall into two classes: (1) {\\it Oxygen and nitrogen, with IP $>$ IP(H), $D_0$(HX) $<$ $D_0$(H$_2$) and $D_0$(HX$^+$) $>$ $D_0$(H$_2$).} These elements are predominantly neutral in the diffuse ISM. Their atoms do not react with H$_2$ except at elevated temperatures that can drive endothermic reactions, and their chemistry is driven by cosmic ray ionization. (2) {\\it Carbon, silicon, sulfur, and phosphorus, with IP $<$ IP(H), $D_0$(HX) $<$ $D_0$(H$_2$) and $D_0$(HX$^+$) $<$ $D_0$(H$_2$).} These elements are predominantly ionized in the diffuse ISM. Their ions do not react with H$_2$ except at elevated temperatures that can drive endothermic reactions. Chlorine and fluorine form two additional classes with one member each: (3) {\\it Fluorine, with IP $>$ IP(H) and $D_0$(HX) $>$ $D_0$(H$_2$).} Fluorine is predominantly neutral in the diffuse ISM, and can react exothermically with H$_2$ to form HF. (4) {\\it Chlorine, with IP $<$ IP(H), $D_0$(HX) $<$ $D_0$(H$_2$) and $D_0$(HX$^+$) $>$ $D_0$(H$_2$).} Chlorine is predominantly ionized in atomic clouds. Cl$^+$ can react exothermically with H$_2$ to form HCl$^+$. The chemistry of interstellar fluorine has been considered in a previous study by Neufeld, Wolfire \\& Schilke (2005; hereafter NWS05), while that of chlorine has been discussed in several studies, most recently those of Schilke, Wang \\& Phillips (1995) and Amin (1996). In the present paper, we extend and update these previous investigations to account for recent developments in experimental and theoretical studies of relevant chemical processes. Our study is also motivated by the prospects for upcoming observations of halogen-containing molecules, particularly with the {\\it Herschel Space Observatory}. ", "conclusions": "As expected, the results shown in Figures 3 -- 7 indicate qualitatively-different behavior for the F- and Cl-bearing species. The results for F-bearing molecules are qualitatively the same as those obtained previously by NWS05: once H$_2$ becomes abundant, HF becomes the dominant reservoir of gas-phase fluorine, with CF$^+$ the only other significant molecule containing fluorine. Because of the lower dissociative recombination rate assumed for CF$^+$ in the present work, the predicted CF$^+$ abundances are somewhat larger. The chemistry for Cl-bearing molecules is somewhat more complex, by contrast, and the computer-generated network diagrams shown in Figure 9, together with Figures 10 - 13 in the online materials, help elucidate the key pathways. \\subsection{Identification of key chemical pathways} In Figures 9 -- 13, circles represent the key Cl-bearing species, with a color scale indicating their relative abundance. The color map extends from cyan to blue to purple to red (highest value), with grey circles denoting species that account for less than 10$^{-8}$ of the gas-phase chlorine abundance. Arrows indicate the key reactions linking the various species: again, the colors of the arrows indicate the reaction rate per unit volume. The results apply to a one-sided model with $n_{\\rm H} = 10^4\\,\\rm cm^{-3}$ and $\\chi_{UV} = 10^4$, and the different figures show the behavior at various depths into the cloud. The depths were chosen to illustrate several different regimes. 1) $A_V = 0$: At the cloud surface, there is no shielding and the abundance of molecules is small. The Cl$^+$/Cl ratio is $\\sim 10^5$, being determined by the balance between photoionization and recombination. For the latter process, radiative recombination and charge transfer recombination with H -- although endothermic by $\\sim$~0.6 eV -- are of roughly equal importance. 2) $A_V= 0.8$: Here, H$_2$ self-shielding has greatly increased the H$_2$ abundance. Reaction with H$_2$ is now the dominant destruction mechanism for Cl$^+$, resulting in the formation of HCl$^+$. However, the e/H$_2$ abundance ratio is still sufficient for dissociative recombination to dominate the destruction of HCl$^+$, although the competing reaction with H$_2$ produces a small amount of H$_2$Cl$^+$. The dissociative recombination of HCl$^+$ leads to atomic Cl, which becomes the dominant gas-phase reservoir of chlorine. Beyond $A_V \\sim 0.8$, the abundances of the HCl$^+$ and H$_2$Cl$^+$ ions start to fall, along with that of $\\rm Cl^+$. 3) $A_V = 2.0$: The e/H$_2$ abundance ratio has now dropped sufficiently that HCl$^+$ reacts primarily with H$_2$, rather than electrons, forming H$_2$Cl$^+$. HCl is produced by dissociative recombination of H$_2$Cl$^+$, although the dominant formation process for HCl is the slightly endothermic reaction of H$_2$ and Cl; this reaction is still possible at the $\\sim 200$~K temperature at this point within the PDR. At $A_V = 2.0$, Cl and HCl account respectively for $\\sim 0.1 \\%$ and $\\sim 99.9 \\%$ of elemental chlorine in the gas phase. 4) $A_V = 4.0$: The gas temperature drops below the level at which HCl can be formed by reaction of H$_2$ with Cl, and the HCl abundance is lower than it was at $A_V = 2.0$. The Cl$^+$/Cl ratio is now only $10^{-8}$, the photoionization rate of Cl having dropped dramatically. Thus the formation of HCl$^+$ is dominated by reaction of H$_3^+$ with Cl, not H$_2$ with Cl$^+$. The H$_3^+$ molecular ion is produced by the effects of cosmic rays; these ionize H$_2$, forming H$_2^+$, which then transfers a proton to H$_2$ to form H$_3^+$. 5) $A_V = 10.0$: As at $A_V=4.0$, HCl$^+$ is produced primarily by reaction of H$_3^+$ with Cl, and reaction of HCl$^+$ with H$_2$ then forms H$_2$Cl$^+$. At this depth, the electron fraction is so low that dissociative recombination of H$_2$Cl$^+$ is no longer dominant. H$_2$Cl$^+$ transfers a proton to a neutral species of higher proton affinity than HCl (e.g. CO or H$_2$O; this process is denoted by the arrow labeled ''N''), yielding HCl. At this point, the external UV is almost completely attenuated, and the HCl destruction rate is very small. HCl now accounts for several $\\times 10\\%$ of the gas-phase chlorine abundance, with atomic chlorine accounting for essentially all of the remainder. Our results in this regime are in good agreement with those obtained previously by Schilke et al\\ (1995). H$_2$Cl$^+$ and CCl$^+$ are also present, but their abundances are just a few $\\times 0.01\\%$ of gas-phase elemental chlorine. HCl reacts with H$_3^+$ to form H$_2$Cl$^+$, although this rapidly leads to reformation of HCl by proton transfer to other neutral species. Several other destruction processes are of comparable importance for HCl destruction: cosmic ray induced photodissociation, and reaction with the positive ions C$^+$ and He$^+$. In this regime, the rates of HCl formation and destruction all scale linearly with the cosmic-ray ionization rate, with the result that the HCl abundance is almost independent of the latter. Although the diagrams shown in Figure 9 -- 13 apply to just a single model, similar regimes exist for other values of $\\chi_{UV}$ and $n_{\\rm H}$. Although the relevant regions move in or out (to larger or smaller $A_V$) as the assumed $\\chi_{UV} / n_{\\rm H}$ ratio is increased or decreased, the same qualitative behavior is observed. In examining similar diagrams for the entire set of models, we have not identified any important reaction pathway that is not apparent in Figures 9 -- 13. \\subsection{Observational implications} To date, HF, CF$^+$, and HCl are the only halogen-bearing molecules to have been detected in the interstellar medium. In regard to HF, our results are identical to those presented in NWS05: we predict HF to be the dominant gas-phase reservoir of fluorine within both diffuse and dense molecular clouds; we expect the {\\it Herschel Space Observatory} to detect widespread absorption in the HF $J=1-0$ transition. However, the abundances we predict for CF$^+$ lie a factor $\\sim 3$ above the prediction of NWS05, a direct consequence of the smaller dissociation rate adopted for CF$^+$ in the present study. As noted by Neufeld et al. (2006), the CF$^+$ column densities inferred from observations of the Orion Bar were already a factor of $\\sim 4$ below the predictions of NWS05, and the discrepancy between theory and observation is now increased to more than an order of magnitude. This disagreement may indicate that we have overestimated the rate coefficient for reaction of C$^+$ and HF. We are not aware of any laboratory studies of this reaction; the rate we adopt is simply the capture rate, and thus an upper limit on the true reaction rate. Laboratory measurements of this key reaction would be very desirable. HCl has been observed in both diffuse and dense molecular clouds. In diffuse clouds, ultraviolet absorption studies have led to upper limits, and in one case, a tentative detection toward $\\zeta$ Oph (Federman et al.\\ 1995). The latter result, obtained using the Goddard High Resolution Spectrograph on HST, yielded an HCl column density of $2.7 \\pm 1.0 (1\\sigma) \\times 10^{11} \\rm cm^{-2}$. Given an atomic chlorine column density of $3.0 \\pm 1.0 (1\\sigma) \\times 10^{14} \\rm cm^{-2}$ for this sight-line (Federman et al.\\ 1995), the corresponding $N({\\rm HCl})/N({\\rm Cl})$ ratio is $9 \\times 10^{-4}$ (uncertain by a factor $\\sim 2$), a value in excellent agreement with predictions of the diffuse cloud models presented by vDB86. Very similar results are obtained from our updated treatment of Cl chemistry. In Figure 14, we show predictions from a series of models with $\\chi_{UV}$ ranging from $10^{-1}$ to 10$^2$. All models apply to a slab with parameters appropriate to $\\zeta$~Oph, viz.\\ density $n_H = 10^{2.5} \\rm \\, cm^{-3}$ and total visual extinction of 0.8 mag. The horizontal axis shows the H$_2$ column density, for which the measured value is $4.2 \\times 10^{20}$ along this sight-line, while the vertical axis shows $N({\\rm HCl})/N({\\rm Cl})$ (red curve), $N({\\rm Cl^+})/N({\\rm Cl})$ (orange) and $N({\\rm HCl^+})/N({\\rm Cl})$ (blue). Squares located along the curves show the results for $\\chi_{UV} = 10^{-1}$ to 10$^2$ from right to left. The H$_2$ column density along the $\\zeta$~Oph sight-line requires a $\\chi_{UV} \\sim 10^{0.5}$, in agreement with previous studies, and the predicted $N({\\rm HCl})/N({\\rm Cl})$ is in good agreement with the observations. The almost exact agreement between the present model and that of vDB86 in regard to the predicted HCl abundance appears to result from the fortuitious cancellation of two changes in the photochemical network: the photodissociation rate for HCl and the photoionization rate for Cl are both increased by a factor $\\sim 2$. As in the vDB86 models, the predicted Cl$^+$ column density lies substantially below the observations. Federman et al.\\ (1995) attributed this discrepancy to the presence of a significant Cl$^+$ column density within HII regions along the sight-line. One important caveat must be noted. The results shown in Figure 14 were obtained by assuming a branching ratio of only 10$\\%$ for the production of HCl following dissociative recombination of H$_2$Cl$^+$. This value was initially invoked primarily to {\\it explain} the relative low HCl abundance derived from upper limits obtained with the {\\it Copernicus} satellite. Models that we obtained for other branching ratios (0, 0.3 and 1.0) indicate that the HCl column density is linearly proportional to the branching ratio for values greater than 0.1 With a branching ratio of zero (i.e.\\ with no production of HCl from dissociative recombination of H$_2$Cl$^+$), a non-zero but negligible HCl column density ($N({\\rm HCl})/N({\\rm Cl}) \\sim 10^{-6}$) results from the endothermic reaction of H$_2$ with Cl. In dense clouds, HCl has been unequivocally detected in several sources by means of submillimeter observations of the $J=1-0$ emission line. First detected using the Kuiper Airborne Observatory toward Orion (Blake, Keene \\& Phillips 1985) and then Sgr B2 (Zmuidzinas et al.\\ 1995), the $J=1-0$ transition has also been observed using the ground-based Caltech Submillimeter Observatory (CSO) under good atmospheric conditions. CSO observations have led to additional detections toward Orion A, Mon R2 (Salez, Frerking \\& Langer 1996) and several other sources (Phillips et al.\\ 2009), as well as mapping observations toward OMC-1 (Schilke, Phillips \\& Wang 1995). Typical column densities derived for these dense clouds lie in the few $\\rm \\times 10^{13}$ to few $\\times 10^{14} \\, \\rm cm^{-2}$ range, in reasonable agreement with the predictions of our model for regions at high density exposed to strong UV radiation (Fig.\\ 5). As noted in the observational studies cited above, depletion plays an important role in limiting the fractional abundance of HCl. Our study identifies two additional Cl-bearing species that are potentially detectable: the molecular ions HCl$^+$ and H$_2$Cl$^+$. These ions are isoelectronic with OH and H$_2$S respectively. They are most abundant near cloud surfaces, where the photoionization rate for HCl is highest, and show column densities that are an increasing function of $\\chi_{UV} / n_{\\rm H}$. Thus, PDRs subject to strong UV irradiation present attractive targets for searches for HCl$^+$ and H$_2$Cl$^+$ at millimeter and submillimeter wavelengths. For example, in a PDR with $n_{\\rm H} = 10^4\\,\\rm cm^{-3}$ and $\\chi_{UV} = 10^4$, column densities of $5.7$ and $2.6\\times 10^{11} \\, \\rm cm^{-2}$ are predicted for HCl$^+$ and H$_2$Cl$^+$. These values apply to sight-lines perpendicular to the illuminated surface; in edge-on PDRs, they can be significantly enhanced by limb-brightening. The rotational spectrum of HCl$^+$ has been measured using laser magnetic spectroscopy (Lubic et al.\\ 1989). The lowest-lying rotational transition, the $^2\\Pi_{3/2} J = 5/2 \\rightarrow 3/2$ multiplet near 207.6$\\mu$m, lies in a wavelength range accessible to the HIFI instrument on the {\\it Herschel Space Observatory}. The rotational spectrum of H$_2$Cl$^+$ has also been the subject of a laboratory study at high spectral resolution; here, the measurements of Araki et al.\\ (2001) indicate that the lowest-lying rotational line of ortho-H$_2$Cl$^+$, the $1_{10}-1_{01}$ transition, lie near 189.2 and 188.4 GHz respectively for the $\\rm H_2^{35}Cl^+$ and $\\rm H_2^{37}Cl^+$ isotopologues. (These transitions are split into 6 and 4 hyperfine components, covering $\\sim 56$ and $14$~MHz respectively.) This spectral region lies in the wing of the strong 183 GHz telluric water absorption line, but is nevertheless observable from ground-based observatories under favorable -- but by no means unusual -- atmospheric conditions. In one of the earliest studies of interstellar chlorine chemistry, the possibility of detecting HCl$^+$ in {\\it diffuse} molecular clouds was discussed by Jura (1974), who considered the $A^2\\Sigma^+ \\leftarrow X ^2\\Pi$ band that is accessible in the near-ultraviolet region. Given an oscillator strength of $4.15 \\times 10^{-4}$ for the strongest vibrational band (Pradhan, Kirby \\& Dalgarno 1991), and given the abundances predicted in Figure 17, our results for $\\zeta$~Oph would imply equivalent widths of only $\\sim 4 \\times 10^{-6}\\,\\AA$, well below the limit of detectability. Finally, we note that, unlike CF$^+$, the CCl$^+$ ion is typically predicted to have an extremely low abundance. Although CCl$^+$ is produced rapidly by reaction of HCl and C$^+$, these two species show very little overlap: the HCl abundance is small in the surface layers where C$^+$ is abundant, unless the temperature is increased by shock heating, and the C$^+$ abundance is small in the deep cloud interiors where HCl is relatively abundant. Enhanced CCl$^+$ abundances are possible when shock heating is present along with UV irradiation." }, "0910/0910.3922_arXiv.txt": { "abstract": "Microlensing and occultation are generally studied in the geometric optics limit. However, diffraction may be important when recently discovered Kuiper-Belt objects (KBOs) occult distant stars. In particular the effects of diffraction become more important as the wavelength of the observation and the distance to the KBO increase. For sufficiently distant and massive KBOs or Oort cloud objects not only is diffraction important but so is gravitational lensing. For an object similar to Eris but located in the Oort cloud, the signature of gravitational lensing would be detected easily during an occultation and would give constraints on the mass and radius of the object. ", "introduction": "\\citet{Bailey:1976p1677} first argued that small bodies in the distant solar system could be detected through stellar occultations. \\citet{1987AJ.....93.1549R} developed the treatment of occultation by irregular bodies including diffraction. More recently with the discovery of the population of Kuiper Belt objects \\citep{Kuiper:1951p1773,Jewitt:1999p1844}, several research groups have begun searching for more distant and smaller bodies through occultations \\citep{Roques:2000p1661,2008AJ....135.1039B,Roques:2009p1666}. Because Kuiper Belt objects (KBOs) typically subtend small angles, the diffraction of radiation around the objects may be important even at visible wavelengths and naturally more important at longer wavelengths \\citep{Roques:6p1676}. The discovery of more distant and more massive KBOs such as Eris \\citep{2005ApJ...635L..97B} begs the question of whether gravitational lensing of background stars by large KBOs and objects in the Oort cloud \\citep{Oort:1950p1427} is important. \\citet{2002astro.ph..9545C} argued that for distant massive KBOs and Oort cloud objects lensing may be important; furthermore, \\citet{2005ApJ...635..711G} argued that GAIA could measure the astrometric displacement from microlensing by an planet more massive than a few Jupiters within $10^4$~AU regardless of its location on the sky. The technique in this paper probes much lower mass objects, but also exploits lensing to provide constraints on the properties of the asteroid. Generally the diffractive effects of microlensing are neglected because the variation in the time delays across the lens is usually much larger than the coherence time of the observation, $1/\\Delta \\nu$ where $\\Delta \\nu$ is the bandwidth of the observation. However, near a caustic crossing, diffraction may be important as argued by \\citet{1995ApJ...455..443J} to account for rapid variations in the light from Q2237+0305 (The Einstein Cross). Typically the differential time delay highly magnified images for a point lens is about $2 GM/c^3$, the crossing time over the Schwarzschild radius of the lens; consequently, for the diffractive effects of lensing to be observable the Schwarzschild radius of the lens should be comparable or larger than the wavelength of the radiation. The largest of the Kuiper Belt objects, Eris, has a Schwarzschild radius $R_S=2GM/c^2 \\approx 25 \\mu$m. Therefore, quite naturally for observations of large KBOs diffractive microlensing may be important for observations in the near and mid-infrared whenever the gravitational effects are important. This paper focuses on just this regime. The first section, \\S\\ref{sec:diffr-micr}, outlines microlensing in the diffractive regime and generalizes the earlier results to include occultation. This yields a expression for the transmission that is nearly identical to the unlensed result. The next section, \\S\\ref{sec:results}, outlines the types of objects for which diffractive lensing may be important, examines several interesting cases and connects the diffraction patterns to the geometric limit (\\S\\ref{sec:geometric-optics}) in the limit where the lensing is weak. The final section (\\S\\ref{sec:conclusions}) outlines how diffractive lensing could constrain the properties of known objects and speculates on the probability of such lensing events. ", "conclusions": "\\label{sec:conclusions} Microlensing is important for large asteroids (similar to Eris) in the Oort cloud and beyond; furthermore, in the near infrared and red-ward diffraction is important to understand the light curves from the combined microlensing and occultation of background stars by such objects. The effects of microlensing are not covariant with variations in the velocity of source, lens and observer nor with variations in the impact parameter; therefore, observations of diffractive lensing combined with a measured distance to the asteroid constrain the mass and radius of the asteroid, or equivalently with an assumed value of the density of the asteroid, such observations would yield a mass, radius and distance. Observations of diffractive microlensing may provide the only way of estimating the masses of such objects unless they have a satellite as Eris does \\citep{2007Sci...316.1585B}. The angular size of even a large asteroid such as Eris is extremely small at the distance of the Oort cloud, so one would expect that the chance for an occultation would be tiny. Over the course of a year, the asteroid will sweep out a region of the sky. The chance of detection each year is proportional to this area. The semimajor axis of the annual parallactic ellipse on the sky ranges from two to twenty arcseconds for a distance of $10^5$ and $10^4$~AU respectively. The area of sky covered each year is given by the product of the arclength around the parallactic ellipse and angular diameter of the asteroid about 0.33~mas for an asteroid of radius 1200~km at $10^4$~AU; therefore, on average, at the nearer distance the asteroid would cover a region of sky 0.33 mas by 100 arcseconds or about 0.033 square arcseconds or a fraction $6 \\times 10^{-14}$ of the entire sky each year. The solid angle covered decreases inversely with the distance squared until microlensing starts to dominate, subsequently the area is proportional to $d^{-3/2}$. It is safe to assume that the asteroid has sufficient velocity perpendicular to the Earth's orbital motion that it moves much more that one diameter per year (much greater than 8.2 cm/s); therefore, the asteroid will cover a new region each year. The asteroid itself may have a large proper motion that could increase the area further --- this motion is neglected in this analysis. Any star within this area of sky will pass within the region $v7.9$ at cluster masses $M_c> 10^5\\,m_\\odot$. At younger ages, it is still present even in massive clusters, and for $M_c \\leqslant 10^4\\,m_\\odot$ it is larger than 0.1 mag in $(B$$-$$V)$. Only for very massive clusters ($M_c>10^6\\,m_\\odot$) with ages $\\log t< 7.5$ is the offset small (of the order of 0.04 mag) and smaller than the typical observational error of colours of extragalactic clusters.} {} ", "introduction": "\\label{sec:intro} Using data on accurate and homogeneous spatio-kinematic-photometric membership for 650 Galactic open clusters, we previously computed their integrated magnitudes in $B,V,J,H$, and $K_s$-passbands \\citep{intpar}. The magnitudes are based on accurate and uniform data from the \\ascc catalogue and were computed by adding the individual luminosities of the most secure cluster members. In contrast to previous lists of integrated magnitudes \\citep[e.g., those of][]{battin94, lata02} of Galactic star clusters, our data provide an independent and more importantly uniform dataset. Hence, our integrated magnitudes can and should be used as benchmarks for studies of the populations of extragalactic star clusters. \\begin{figure}[t] \\resizebox{\\hsize}{36mm}{ \\includegraphics[clip]{13270f1a.ps} \\includegraphics[clip]{13270f1b.ps} }\\\\ \\resizebox{\\hsize}{36mm}{ \\includegraphics[clip]{13270f1c.ps} \\includegraphics[clip]{13270f1d.ps} } \\caption{Observed colours of 650 Galactic open clusters compared to theoretical colours computed from standard SSP models. The upper row is for $B$$-$$V$, the bottom one for $J$$-$$K_s$. The left column shows ``integrated colour $vs. \\log t\\,$'' diagrams, the right one the distributions of colour indices. Dots show Galactic open clusters from our sample. Black dots are young clusters used for constructing the colour distribution, grey dots the older clusters. The curves are the SSP model tracks (the solid one is from Starburst99, the dashed one is from GALEV).The observed colour distributions are shown as the filled cyan histogram, the GALEV model distribution as the hatched red one. } \\label{sev_fig} \\end{figure} Based on these uniform data, we found disagreement between the observed colours of our cluster sample and theoretical colours derived from simple stellar population (SSP) models. The comparison of the \\citet{intpar} data with present-day SSP models such as GALEV \\citep{galev03} and Starburst99 \\citep{sb99_05} illustrated a substantial offset of the order of $0.2-0.3$ mag in $(B$$-$$V)$ of the model colours with respect to the observations for clusters younger than $\\log t \\lesssim 8.5$. This is both remarkable and important, since most data such as age and mass collected nowadays for the bulk of extragalactic clusters, are derived by comparing their integrated light with the predictions of SSP models \\citep[see e.g.,][]{bikea03}. The models are produced by evolutionary synthesis codes simulating the extensive and complicated stellar populations of galaxies. However, they are also believed to be appropriate for describing cluster populations more than five orders of magnitude less massive (and apparently far simpler). Although star clusters may be influenced by effects that are irrelevant to galaxies (e.g., low number statistics, because of the discreteness of the real IMF, or dynamical mass-loss by evaporation of stars), the validity of the SSP approach for star clusters can be tested from general considerations \\citep{cervino04}. In contrast, in this paper we apply the issue of the discreteness of the IMF to the specific case of Galactic open clusters. A colour offset of this kind was found previously by \\citet{lata02}, in the colour indices $U$$-$$B$, $B$$-$$V$, $V$$-$$R$, and $V$$-$$I$. However, they accept that only the offsets in $V$$-$$R$ and $V$$-$$I$ represent significant deviation of their observations from the model predictions. We suspect that the discreteness of the IMF in open clusters may explain these discrepancies. To study this issue in a more systematic way, we constructed our own SSP model that takes into account the effect of low number statistics (discreteness), which is a common problem for open star clusters. We cross-checked a continuous version of our code with GALEV and found good agreement. The discreteness of the IMF itself is a natural assumption for any stellar ensemble consisting of individual objects. However, when one considers vast stellar populations (e.g., galaxies) that densely populate the entire colour-magnitude diagram, one can adopt a continuous IMF, which is more convenient for different technical reasons. For small populations (stellar clusters), the assumption of a discrete IMF is clearly appropriate.. This letter describes the solution of a restricted problem, namely the influence of the discrete-IMF approach on the colours of young clusters. The full scope of results related to the IMF-discreteness effect on star clusters will be reported elsewhere. ", "conclusions": "\\label{sec:conc} Observations of local Galactic open clusters have enabled us to measure their integrated magnitudes, masses, ages, and reddening. Using this data set, we have found a remarkable discrepancy between the observed colours and the predictions of SSP models. The main reason for this disagreement is the neglect of the assumption of IMF-discreteness. When this effect is taken into account, the model agrees adequately with the observations and is even able to explain the large colour spread observed in the empirical colour-age relation in a natural way. In which conditions is the effect of discreteness relevant? Since it is a consequence of low-number statistics, it depends on the sparseness of the stellar population in the upper CMD of a cluster where the bulk of light is emitted. The density of the population depends primarily on cluster mass, and in addition on cluster age, on [strange at first glance] the lower mass limit of the stars formed, and to some degree on the slope of the IMF. In the context of this letter, other factors are not important. According to our investigation, the \\textit{systematic} offset between the continuous- and discrete-IMF colours diminishes substantially but not completely at $\\log t >7.9$, at cluster masses $M_c> 10^5\\,m_\\odot$. At younger ages, it remains present even in massive clusters, and for $M_c \\leqslant 10^4$ it is larger than 0.1 mag in $(B$$-$$V)$. Only for very massive clusters ($M_c>10^6\\,m_\\odot$) and young ages ($\\log t< 7.5$), the offset falls below a typical observational error. We note in passing that the effect is stronger for redder passbands. These findings are in good agreement with the theoretical forecast of \\citet{cervino04}. The immediate consequence of the application of a continuous-IMF approach to the SSP models of stellar clusters is a systematic underestimate of reddening. Other problems affect the age and mass determination based on evolutionary grids of these models (at least for masses below $10^6\\,m_\\odot$). Because of the rather flat colour-age relation of young clusters, a systematic error of the order of 0.1 mag produces an error in cluster age of about one order of magnitude. On the other hand, there is a danger that the existence of a large number of ``main-sequence'' (i.e., blue) clusters could be interpreted as evidence of a recent burst of star formation. We can exclude such a burst having occured in our Galactic open clusters even if they are blue. Finally, the technique of the parameter determination should be changed to incorporate the cluster flux variations caused by the discrete nature of the upper IMF." }, "0910/0910.5609_arXiv.txt": { "abstract": "}[2]{{\\footnotesize\\begin{center}ABSTRACT\\end{center} \\vspace{1mm}\\par#1\\par \\noindent {~}{\\it #2}}} \\newcommand{\\TabCap}[2]{\\begin{center}\\parbox[t]{#1}{\\begin{center} \\small {\\spaceskip 2pt plus 1pt minus 1pt T a b l e} \\refstepcounter{table}\\thetable \\\\[2mm] \\footnotesize #2 \\end{center}}\\end{center}} \\newcommand{\\TableSep}[2]{\\begin{table}[p]\\vspace{#1} \\TabCap{#2}\\end{table}} \\newcommand{\\FigCap}[1]{\\footnotesize\\par\\noindent Fig.\\ % \\refstepcounter{figure}\\thefigure. #1\\par} \\newcommand{\\TableFont}{\\footnotesize} \\newcommand{\\TableFontIt}{\\ttit} \\newcommand{\\SetTableFont}[1]{\\renewcommand{\\TableFont}{#1}} \\newcommand{\\MakeTable}[4]{\\begin{table}[htb]\\TabCap{#2}{#3} \\begin{center} \\TableFont \\begin{tabular}{#1} #4 \\end{tabular}\\end{center}\\end{table}} \\newcommand{\\MakeTableSep}[4]{\\begin{table}[p]\\TabCap{#2}{#3} \\begin{center} \\TableFont \\begin{tabular}{#1} #4 \\end{tabular}\\end{center}\\end{table}} \\newenvironment{references}% { \\footnotesize \\frenchspacing \\renewcommand{\\thesection}{} \\renewcommand{\\in}{{\\rm in }} \\renewcommand{\\AA}{Astron.\\ Astrophys.} \\newcommand{\\AAS}{Astron.~Astrophys.~Suppl.~Ser.} \\newcommand{\\ApJ}{Astrophys.\\ J.} \\newcommand{\\ApJS}{Astrophys.\\ J.~Suppl.~Ser.} \\newcommand{\\ApJL}{Astrophys.\\ J.~Letters} \\newcommand{\\AJ}{Astron.\\ J.} \\newcommand{\\IBVS}{IBVS} \\newcommand{\\PASP}{P.A.S.P.} \\newcommand{\\Acta}{Acta Astron.} \\newcommand{\\MNRAS}{MNRAS} \\renewcommand{\\and}{{\\rm and }} {We report the analysis of a binary blue straggler in NGC 6752 with a short orbital period of 0.315 d and a W UMA-type light curve. We use photometric data spanning 13 years to place limits on the mass ratio ($0.15\\lesssim q \\lesssim 0.35$), luminosity ratio ($L_{1}/L_{2}\\approx 4.0$) and the ratio of the radii of the components ($r_{1}/r_{2}\\approx 2.0$). The effective temperatures of the components are nearly identical, and the system is detached or semi-detached (in the latter case the component filling its Roche lobe is the secondary). Such a configuration is unusual given the shortness of the orbital period, and it must have resulted from substantial mass exchange. We suggest that some secondaries of W UMa-type stars, normally regarded as main sequence objects which fill their Roche lobes to different degrees, in fact may be shell-burning cores of originally more massive components. } {\\bf Key words:} {\\it blue stragglers -- binaries: eclipsing -- globular clusters: individual (NGC~6752)} ", "introduction": "\\label{sect:intro} Blue stragglers (BSs) occupy the extension of the main sequence above the turnoff point in color-magnitude diagrams of parent stellar populations. Since their detection in M3 (Sandage 1953) BSs have been identified in dozens of globular and open clusters. Two main mechanisms responsible for the formation of these objects have been advocated -- mass transfer in close binaries (McCrea 1964), and mergers resulting from direct collisions between stars in dense environments (Benz \\& Hills 1987). Knigge et al. (2009) found that the number of BSs belonging to a given globular cluster (GC) scales nearly linearly with the core mass of that cluster but is nearly independent of the core radius. This implies that the main channel leading to the formation of BSs is binarity. Moreover, Perets \\& Fabrycky (2009) argue that the progenitors of BSs form in primordial triple stars through the Kozai mechanism (Kozai 1962). This is in line with the observational evidence that most (possibly all) close binaries are formed in triple systems (Tokovinin et al. 2006). The sample of candidate BSs in GCs includes several variables with light curves typical of contact binaries. The discovery of two W~UMa type systems in NGC~5466 (Mateo et al. 1990) was followed by the detection of further such systems in other clusters, mostly by the OGLE and CASE groups (see Rucinski 2000). Surprisingly, until now not a single mass determination or even a reliable light curve solution has been obtained for a GC contact binary. In fact, masses resulting from the analysis of light and radial velocity curves have been derived for just two binary BSs from GCs (Kaluzny et al. 2007a, 2007b). Some attempts have been made to estimate masses of apparently single BSs in GCs based on the modeling of spectra (\\eg De Marco et al. 2005) or pulsation periods of SX~Phe variables (Gilliland 1998). However, the actual accuracy of these determinations is hard to estimate. The eclipsing binary V8-NGC~6752 (hereafter V8) was discovered by Thompson et al. (1999) during a photometric survey of the cluster field. The system has a W~UMa-type light curve, and an orbital period of 0.31~d. Our new photometry (Kaluzny \\& Thompson 2009) revealed that the secondary eclipse of V8 is total. As it was first discussed by Mochnacki \\& Doughty (1972), light curves of contact binaries showing total eclipses can be uniquely solved, yielding the mass ratio among other parameters. This is in contrast to contact systems with partial eclipses whose light curves usually cannot be reliably solved without spectroscopic information about the mass ratio.\\footnote{Except for relatively rare systems with mass ratios close to unity.} In short, the totality of the primary eclipse is sufficient to determine reliable apparent magnitudes of the individual components. If V8 is indeed a contact binary belonging to the cluster then one can estimate its absolute parameters based solely on the photometric information, which together with the known distance to the cluster would yield the absolute magnitudes of its components. Then, based on effective temperatures estimated from colors, one can derive absolute radii of both stars. Knowing absolute radii and relative radii (the latter from the light curve solution) the absolute size of the orbit can then be derived, yielding the total mass of the system from Kepler's law. Given the paucity of reliable determinations of masses for BSs in GCs we decided to analyze the light curves of V8 hoping to estimate its parameters. To our surprise it turned out that the binary is most likely a detached system with a mass ratio $q<0.35$. This is an unexpected result given the short orbital period of the system and the similar effective temperatures of its components. ", "conclusions": "\\label{sect:summary} Our analysis indicates that, despite its short orbital period and W~UMa-type light curve, the binary V8 is a detached or semidetached system. This is an unexpected finding, as non-contact systems are extremely rare among non-degenerate binaries with periods $P\\leq 0.45$~d (Hilditch et al. 1988). To the best of our knowledge the only known systems of this kind are V361~Lyr (Hilditch et al. 1997; Kaluzny 1991; $P=0.30~d$) and OGLE-BW3-V38 (Maceroni \\& Rucinski 1997; $P=0.20$~d). The first of these is a semi-detached system with the more massive component filling its Roche lobe, while the latter is probably a detached system composed of two M dwarfs. Another possible non-contact binary of this kind is W~Crv (Rucinski \\& Lu 2000), however its actual configuration is poorly constrained. We note that V8 cannot be considered as an example of a system caught in the brief contact-break predicted by ``the thermal relaxation oscillations theory\" (Lucy \\& Wilson 1979). In such a system the {\\em primary} of V8 would have to fill its Roche lobe, and the components should have very different effective temperatures (the secondary being cooler). This is certainly excluded. The stability and symmetry of the light curve of V8 together with the apparent constancy of the orbital period favors a detached configuration. Despite the totality of one of the eclipses the photometric analysis does not allow an accurate determination of the mass ratio of the system. However, it strongly favors the range $0.15\\lesssim q \\lesssim 0.35$. The components differ significantly in size ($r_{1}/r_{2}\\approx 2$), yet they have very similar effective temperatures. Figure 5 shows the location of the whole system and each of its components separately on the color-magnitude diagram of NGC~6752. When calculating the individual magnitudes of the components we adopted a luminosity ratio of 3.97, the value implied by the solution with $q=0.20$ for radiative envelopes. Note, however, that the derived value of $L_{1}/L_{2}$ depends rather weakly on the assumed $q$ (see Tables 2 and 3). It is worth noting that the primary of V8 is located right at the extension of the unevolved main sequence of the cluster. This gives one more argument for the cluster membership of the binary, and suggests that the primary has the properties of an undisturbed main-sequence star. In contrast, the secondary is located far to the blue of the cluster main-sequence for its luminosity. Also, the observed luminosity ratio $L_{2}/L_{1}\\approx 1/4$ is much too high compared to that expected for main-sequence stars with $m_{2}/m_{1} < 0.35$. For stars with $0.43$1 pc) and the molecular mass are of the same order as those of known nearby low mass star formation regions like Chamaeleon~I. CG~12 has been recently reviewed by \\citet{2008hsf2.book..847R}. The median photometric age of the T-Tauri population in CG~12 is $\\sim$4 Myr, but the apparent spread of ages is considerable, from 1 to 20 Myr \\citepalias{getmanetal2008}. The more than 50 T-Tauri stars detected by \\citetalias {getmanetal2008} represent possibly the first generation of stars born in the CG~12 area, and the B and A stars \\citep{williams1977} may be the second generation. The four IRAS point sources and the collimated molecular outflow in the CG~12 cloud show that star formation is presently going on. Because of the great distance and the high extinction towards the cloud core, the limiting magnitudes of the Two Micron All Sky Survey (2MASS) survey allow only the detection of the brightest stars embedded in the cloud in all three \\jhks \\ colours. A deeper \\J,\\H,\\Ks \\ survey of CG~12 than 2MASS is clearly needed to study the possible embedded stellar population. Spitzer NIR IRAC imaging data on CG~12 has recently become public and can also be used to analyse the embedded stars. We report imaging of the core of CG~12 in \\J, \\H, and \\Ks \\ NIR bands with SOFI at the NTT telescope at La Silla. Observations, data reduction, and calibration procedures are described in Sect.~\\ref{sec:observations} and the observational results in Sect~\\ref{sec:results}. The new results are discussed and compared with available data at other wavelengths in Sect.~\\ref{sec:discussion}. The results are summarised and the conclusions drawn in Sect.~\\ref{sec:summary}. ", "conclusions": "\\label{sec:summary} The head of the cometary globule CG~12 has been observed in the \\J, \\H, and \\Ks \\ bands. The new NIR photometry combined with already existing data at other wavelengths allows the analysis of the star formation and the content of the associated stellar cluster in more detail than was possible before. Star formation in CG 12 (within the region imaged in \\jhks \\ with SOFI) is concentrated on and near the two strong \\ceo maxima \\cgs and \\cgnp. Apart from the known members of the associated stellar cluster several new, embedded cluster members were detected as well. Because the SOFI images are much deeper than the available 2MASS survey data it is possible to define more accurately the NIR colours of the already known embedded member stars which were near the 2MASS limit. Apart from the stellar photometry, the SOFI imaging allows the study of the cloud surface brightness due to scattered light. Scattered light permits us to pinpoint the central source of the highly collimated molecular outflow detected by \\citet{white1993} and to detect the probable shadow cast by the circumstellar disk around star \\four. Using NIR \\J,\\H,\\Ks \\ photometry alone is not in all cases sufficient to decide whether some stars are indeed members of the CG~12 stellar cluster. In some cases an associated nebulosity can confirm the membership. For four stars the observed X-ray activity \\citepalias{getmanetal2008} indicates that the stars are likely members. The conclusions are the following: 1. Seven embedded (proto)stars (\\one, \\two, \\three, \\four, \\five, \\six\\ and W77-6b) are identified. The visual extinction of these stars ranges from 10 to more than 30 magnitudes. The stars \\three \\ and \\four \\ were already known members. Stars \\two, \\ \\three, \\ \\four \\ and \\five \\ are also detected in X-rays \\citepalias{getmanetal2008}. 2. Stars \\one, \\four, \\five, \\six, and W77-6b have infrared excess indicating a circumstellar shell or disk. The shadow of the probable disk around star \\four \\ is seen in the cloud surface brightness. 3. Scattered light from the probable central source of the collimated outflow detected by \\citet{white1993} as well as an hourglass-shaped cavity in the parental molecular cloud formed by the outflow are detected. Apart from the hourglass a larger scale cavity cleared by the outflow to the less tenuous part of the molecular cloud is also seen. The stellar source is better resolved in the Spitzer IRAC images at 3.4, 4.5 and 5.8 $\\mum$ \\ than in the SOFI \\H \\ and \\Ks \\ images, where it blends into the bright background nebulosity. {\\bf 4.} The \\citet{williams1977} star W77-6 is a binary with a separation of $\\sim$4\\arcsec. The brighter component, W77-6a is an early A type star with a visual extinction of $\\sim$3 magnitudes. The fainter component, W77-6b, has NIR excess indicating circumstellar material. The W77-6b \\J \\ magnitude changed by 0\\fm6 between two observing nights. The Spitzer NIR imaging suggests that a third source and an associated nebulosity not visible in the \\jhks \\ imaging may be present. 5. HIRES enhanced IRAS 12 and 25\\,$\\mum$ \\ maps combined with IRAS fluxes and the NIR imaging suggest that the IRAS point source 13547--3944 consists of two separate sources. The 12 and 25 $\\mum$ \\ emission originates in the h4636n component and the longer wavelengths from an adjacent, embedded cold cloud source. The 12 $\\mum$ \\ HIRES map of IRAS 13546--3941 resolves into two previously undetected sources coinciding with the binary stars W77-6a/b and \\three/\\four. The HIRES 25 $\\mum$ maximum lies between the 12 $\\mum$ \\ sources near the IRAS 13546--3941 nominal position and the \\ceo maximum and a mm continuum source. The 25\\,$\\mum$\\ detection could be a source embedded in the continuum core. 6. The SOFI data are not deep enough to probe the visual extinction through the two dense cloud cores in the direction of \\cgn and \\cgsp. The observed maximum extinction around \\cgn is $\\sim$20\\umag, and $\\sim$15\\umag \\ around \\cgsp. No background stars are seen through the core centres." }, "0910/0910.3733_arXiv.txt": { "abstract": "% Image compression has been a frequent topic of presentations at ADASS. Compression is often viewed as just a technique to fit more data into a smaller space. Rather, the packing of data -- its ``density'' -- affects every facet of local data handling, long distance data transport, and the end-to-end throughput of workflows. In short, compression is one aspect of proper data structuring. For example, with FITS tile compression the efficient representation of data is combined with an expressive logistical paradigm for its manipulation. \\par A deeper question remains. Not just how best to represent the data, but which data to represent. CCDs are linear devices. What does this mean? One thing it does not mean is that the analog-to-digital conversion of pixels must be stored using linear data numbers (DN). An alternative strategy of using non-linear representations is presented, with one motivation being to magnify the efficiency of numerical compression algorithms such as Rice. ", "introduction": "The Rice compression algorithm (Rice {\\em et\\ al.} 1993) is particularly familiar to astronomers in combination with the FITS Tiled Image Convention (White {\\em et\\ al.} 2006, Seaman {\\em et\\ al.} 2007). It has been informally paired in the past with non-linear data representations (Nieto-Santisteban {\\em et\\ al.} 1999, Nicula {\\em et\\ al.} 2005). Mention of compression is often followed immediately by the word ``scheme'', as if this branch of computer science were suitable only for black-box heuristics arrived at through a process of trial-and-error. This paper seeks to begin a process of layering the issue of optimal data encoding onto a more formal foundation. As discussed in Pence, Seaman \\& White (2009A,B), the compression achieved for an astronomical image is typically determined almost entirely by its background noise. Any processing that quiets the background will increase the compression ratio. This is the foundation for all lossy compression algorithms -- in effect, lossy compression combines a preprocessing step that tempers an image's noise characteristics with the subsequent application of a lossless algorithm. Janesick (2001) describes a non-linear hardware or software component called a ``square-rooter''. Instead of using a linear analog-to-digital conversion when reading out a CCD, the square-root of these values acts to linearize the Poisson statistics that characterizes these detectors. Such a DN encoding has been used on several space missions to reduce bandwidth requirements for data transport. While technically lossy if the square-root transform occurs after the A/D conversion, it is equally reasonable to consider this a type of lossless encoding since the transform maintains oversampling of the noise at both the high and low end of the dynamic range. Square-root encoding itself acts as a form of compression -- 65,636 linear data levels (for 16-bit pixels) will turn into a scant 256 levels after the square-root. This is not nearly as dramatic when expressed in bits, since mapping 16 bits into 8 bits corresponds to a compression ratio of just $R = 2$. The more important aspect, rather, is to linearize the noise. ", "conclusions": "" }, "0910/0910.4113_arXiv.txt": { "abstract": "We study the spectroscopic properties and environments of red (or passive) spiral galaxies found by the Galaxy Zoo project. By carefully selecting face-on, disk dominated spirals we construct a sample of truly passive disks (\\ie ~they are not dust reddened spirals, nor are they dominated by old stellar populations in a bulge). As such, our red spirals represent an interesting set of possible transition objects between normal blue spiral galaxies and red early types, making up $\\sim6\\%$ of late-type spirals. We use optical images and spectra from SDSS to investigate the physical processes which could have turned these objects red without disturbing their morphology. We find red spirals preferentially in intermediate density regimes. However there are no obvious correlations between red spiral properties and environment suggesting that {\\it environment alone is not sufficient to determine if a galaxy will become a red spiral}. Red spirals are a very small fraction of all spirals at low masses ($M_\\star < 10^{10}$\\msun), but are a significant fraction of the spiral population at large stellar masses showing that {\\it massive galaxies are red independent of morphology}. We confirm that as expected, red spirals have older stellar populations and less recent star formation than the main spiral population. While the presence of spiral arms suggests that major star formation cannot have ceased long ago (not more than a few Gyrs), we show that these are also not recent post-starburst objects (having had no significant star formation in the last Gyr), so {\\it star formation must have ceased gradually}. Intriguingly, red spirals are roughly four times as likely than the normal spiral population to host optically identified Seyfert/LINER (at a given stellar mass and even accounting for low luminosity lines hidden by star formation), with most of the difference coming from objects with LINER-like emission. We also find a curiously large optical bar fraction in the red spirals ($70\\pm5\\%$ verses $27\\pm5\\%$ in blue spirals) suggesting that {\\it the cessation of star formation and bar instabilities in spirals are strongly correlated}. We conclude by discussing the possible origins of these red spirals. We suggest they may represent the very oldest spiral galaxies which have already used up their reserves of gas - probably aided by strangulation or starvation, and perhaps also by the effect of bar instabilities moving material around in the disk. We provide an online table listing our full sample of red spirals along with the normal/blue spirals used for comparison. ", "introduction": "The advent of large galaxy surveys like the Sloan Digital Sky Survey (SDSS) in which photometry (and therefore colours) are readily available for millions of objects has lead to the common use of optical colours to define ``early\" and``late\" type galaxy samples \\citep[e.g.][]{S09,S08,LP07,C07,B06,C05}. This method is particularly favoured since obtaining morphologies for large numbers of galaxies has until recently been impossible. This simplification is justified since it has been shown many times that the majority of galaxies follow a strict colour-morphology relation. For example \\citet{M09} argued that 85\\% of galaxies to z$\\sim$1 are either red, bulge-dominated galaxies or blue, disk dominated galaxies; while \\citet{C06} showed a similar result for 22000 low redshift galaxies (both using automated methods for morphological classification). However the clear correlation between colour and morphology is surprising, given that the colours of galaxies are determined primarily by their stellar content (and therefore their recent star formation history, mostly within the last Gyr) while the morphology is primarily driven by the dynamical history. The clear link between colour and morphology then gives a strong indication that the timescales and processes which drive morphological transformation and the cessation of star formation are strongly related - at least in most cases. In this paper however, we consider a class of object (the red spirals) where the link described above appears to be broken. Since the morphology-density relation was first quantified \\citep{D80} many mechanisms have been proposed for the transformation of blue, star forming, disk galaxies in low density regions of the universe, to red, passive, early type galaxies in clusters. A recent review of many of the proposed mechanisms, and the evidence supporting them, can be found in \\citet{BG06}. Clearly two things must happen for a star forming blue spiral galaxy to turn into a passive red early type. First, star formation must cease (which can indirectly alter the morphology by causing spiral arms and the disk in general to fade, possibly producing an S0 or lenticular from a spiral), and secondly, in order to produce a {\\it bona fide} elliptical the same, or a different process must also dynamically alter the stellar kinematics of the galaxy. The presence of an unusually red or passive (\\ie. non-star forming) population of spiral galaxies in clusters of galaxies was first noticed by \\citet{V76} in the Virgo cluster. Later studies of distant cluster galaxies in {\\it Hubble Space Telescope} (HST) imaging also revealed a significant number of so-called ``passive'' spiral galaxies with a lack of on-going star formation (Couch et al. 1998; Dressler et al. 1999; Poggianti et al. 1999). Passive late-type galaxies were identified at lower redshifts in the outskirts of SDSS clusters by \\citet{G03}, using concentration as a proxy for morphology. Passive spirals in a cluster at $z\\sim0.4$ were studied by \\citet{M06} who found star formation histories from GALEX observations consistent with the shutting down of star formation from strangulation (as described by \\citealt{Bekki02}). Passive spirals have also been revealed in a cluster at $z\\sim 0.1$ in the STAGES survey, using HST morphologies \\cite{W09}, rest frame NUV-optical SEDs \\citep{W05} and 24$\\mu$m data from Spitzer \\citep{G09}. In that series of papers, ``dusty red late-types\" and ``optically passive late-types\" are found to be largely the same thing, with a non-zero (but significantly lowered) star formation rate revealed by the IR data. Red spirals/late types have been studied in several recent papers \\citep{Lee08,D09,HC09,CH09}, as well as \\citet{MR09} who talk about blue passive galaxies (\\ie ~ galaxies with blue colours, but showing no indication of recent star formation in their spectra) which mostly appear to have late type morphologies and have very recently shut down star formation. These might be the progenitors of the red spirals. \\citet{Bundy09} have studied the redshift evolution of red sequence galaxies with disk like components in COSMOS and use it to estimate that as many as 60\\% of spiral galaxies must pass through this phase on the way to the red sequence - making it an important evolutionary step. The Galaxy Zoo project \\citep{L08} revealed the presence of a significant number of visually classified spiral galaxies which are redder than the blue cloud (between 16-28\\% of the total galaxy population depending on environment, \\citealt{B09}). In this paper we study in more detail the physical properties and environments of this population of red spiral galaxies. Galaxies drawn from this population have the morphological appearance of spiral galaxies with a distinct spiral arm structure, but have rest-frame colours which are as red as a typical elliptical galaxy, indicating little or no recent star formation activity. We are studying these objects in order to identify the physical process which is most important in their formation. It is clear that all spiral galaxies can be affected by various physical processes as they evolve -- in this paper we attempt to identify which are most important for red spirals, asking how they are able to shut down star formation while retaining their spiral morphology. A list of possible mechanisms includes processes that depend on environment such as: (1) galaxy-galaxy interactions: in high density regions there is an increased probability of interaction with other galaxies. Most major mergers destroy spiral structure \\citep{TT72} unless they involve very gas rich progenitors \\citep{H09}, but some interactions can be quite gentle (\\eg~ Walker, Mihos \\& Hernquist 1996), for example minor-mergers, tidal interactions etc. (2) Interaction with the cluster itself also occurs and can remove the gas which forms the reservoir for star formation. This can be due to tidal effects \\citep[e.g.][]{Gn03}, or interaction with the hot intercluster gas, either through thermal evaporation \\citep{CS77} or ram pressure stripping \\citep{GG72}. (3) Processes like harrassment \\citep{M99} and starvation or strangulation \\citep{L80,Bekki02} have also been shown to have a significant effect on late type galaxies. Harrassment refers to the heating of gas by many small interactions, while starvation or strangulation refers to the gradual exhaustion of disk gas after the hot halo has been stripped away. These mechanisms both occur at much larger cluster radii (\\ie ~lower densities) than the \"classic\" environmental effects. Internal mechanisms could be more important. For example, (4) the latest semi-analytical models of galaxy formation all invoke feedback from a central massive black hole (or active galactic nuclei; AGN) to explain the most massive red elliptical galaxies \\citep{G04,S05,Sc06,Cr06,Bower06}, although the effect of this process on disk galaxies has been studied less, it still may have some effect (Okamoto, Nemmen \\& Bower 2008). (5) Another culprit could be bar instabilities in spiral galaxies which drive gas inwards (eg. Combes \\& Sanders 1981) and may trigger AGN activity and/or central star formation (eg. Shlosman et al. 2000), perhaps using up the reservoir of gas in the outer disk and making spirals red. (6) Finally red spirals could simply be old spirals which have used up all their gas in normal star formation activities without having any major interactions. In normal spirals, the gas that feeds ongoing star formation comes from infall of matter from a reservoir in the outer halo \\citep{BG06}. As first suggested by Larson, Tinsley \\& Caldwell (1980) and expanded by \\citet{Bekki02} the removal of gas from this outer halo (``strangulation\", or ``starvation\") will cause a gradual cessation of star formation proceeding over several Gyrs. We describe the sample and data along with the selection of the ``red spirals\" in more detail in Section 2. In Section 3 we discuss the stellar populations and star formation history of the red spirals. In Section 4 we discuss their environment and the environmental dependence of star formation. The impact of AGN is considered in Section 5, and bar instabilities are discussed in Section 6. In Section 7 we discuss the plausible mechanisms for formation of the red spirals and future directions which could be taken to distinguish between them. We present a summary and conclusions in Section 8. The adopted cosmological parameters throughout this paper are $\\Omega_{m}$ = 0.3, $\\Omega_{\\Lambda}$ = 0.7 and $H_{0} = 70$ \\kmsMpc. ", "conclusions": "We study the interesting population of red, or passive, spiral galaxies found by the Galaxy Zoo project. We identify from these red spirals a population of intrinsically red true disk dominated spirals by limiting the sample in inclination (to reduce the impact of dust reddening), requiring that spiral arms be visible, and removing bulge dominated systems using the SDSS parameter {\\tt fracdeV}. We compare this sample to blue spirals selected in the same way and find that: \\begin{itemize} \\item{Red spirals are more likely to be more massive, luminous galaxies than blue spirals. They represent an insignificant fraction of the spiral population at masses below $10^{10}$\\msun ~but are significant at the highest masses, showing that {\\it massive galaxies are red regardless of morphology}.} \\item{At the same mass as blue spirals, face-on red spirals do not have larger amounts of dust reddening (as measured by the Balmer decrement), therefore their red colours indicate an ageing stellar population not an increased dust content.} \\item{Red spirals have lower (but not zero) rates of ongoing and recent star formation when compared to blue spirals. This is partly related to their higher average mass, however {\\it at a fixed mass, red spirals still have less recent star formation than blue spirals}.} \\item{As previously observed, red spirals are more common at intermediate local densities (around, or just inside the infall regions of clusters). They are also observed to be more likely than blue spirals to have close neighbours.} \\item{Red spirals in all environments have lower rates of recent and on-going star formation than blue spirals, and there are no significant trends of the star formation rates with environment when spirals are split into red/blue. Clearly the process which creates red spirals is not confined to regions of high galaxy density. So {\\it environment alone is not sufficient to determine whether a galaxy will become a red spiral or not}.} \\item{Red spirals are more than four times more likely to be classified as Seyfert+LINER/composite objects from their optical spectra than blue spirals. This is partly due to the higher masses of red spirals but is still observed when they are compared to a blue spiral sample selected to have the same mass distribution. We find that a small fraction of low luminosity AGN are being revealed as the star formation is turned off in the red spirals, but this is not enough to account for all the difference. Most of the difference comes from an increased fraction of LINER-like emission ($82\\pm12\\%$ of Seyfert/LINERs found in red spirals are LINERs compared to $57\\pm7\\%$ in blue spirals).} \\item{Red spirals have significantly higher bar fractions than blue spirals (70\\% versus 27\\%), suggesting that bar instabilities and the shutting down of star formation in spirals are correlated}. \\end{itemize} We propose three possible origins for the red spiral population studied in this work and suggest the most likely explanation is that a combination of the three accounts for the shutting down of their star formation while they retain their spiral structure: \\begin{enumerate} \\item{Perhaps red spirals are just old spirals which have used up all of their gas. They are found preferentially in intermediate density regions because structures first starts to form at the peaks of the dark matter distribution, but in the centres of clusters spiral morphologies cannot stand up to the environmental disturbances. Red spirals then represent the end stages of spiral evolution irrespective of environment (and in the absence of major mergers) - the spiral version of ``downsizing\".} \\item{Perhaps red spirals are satellite galaxies in massive dark matter halos. In this scenario, they are accreted onto the halo as a normal blue spiral and have experienced either strangulation (where the gas in their outer halos has been gently stripped off, and no further cold gas has been allowed to accrete) or harassment (heating their disk gas and preventing further star formation). Low mass spirals would probably be disrupted in this process and so are not observed as red spirals.} \\item{Perhaps red spirals evolved from normal blue spirals which had bars that were particularly efficient at driving gas inwards. This removed gas from the outer disk and turned the spiral red. If it triggered star formation in the central regions it must have occured more than $\\sim 1$ Gyr ago since red spirals are not post starburst galaxies.} \\end{enumerate} The red spirals in this work probably cannot be the progenitors of S0s as they have a significantly higher bar fraction than in observed in the S0 population. S0s may however be the end product of red spirals with larger bulges than we have studied here. \\paragraph*{ACKNOWLEDGEMENTS} This publication has been made possible by the participation of more than 160,000 volunteers in the Galaxy Zoo project. Their contributions are individually acknowledged at \\texttt{http://www.galaxyzoo.org/Volunteers.aspx}. KLM acknowledges funding from the Peter and Patricia Gruber Foundation as the 2008 Peter and Patricia Gruber Foundation International Astronomical Union Fellow, and from the University of Portsmouth and SEPnet (www.sepnet.ac.uk). Support for the work of MM in Leiden was provided by an Initial Training Network ELIXIR (EarLy unIverse eXploration with nIRspec), grant agreement PITN-GA-2008-214227 (from the European Commission). AKR, MM, HCC, RCN acknowledge financial support from STFC. Support for the work of KS was provided by NASA through Einstein Postdoctoral Fellowship grant number PF9-00069 issued by the Chandra X-ray Observatory Center, which is operated by the Smithsonian Astrophysical Observatory for and on behalf of NASA under contract NAS8-03060. CJL acknowledges support from The Leverhulme Trust and the STFC Science In Society Programme. Funding for the SDSS and SDSS-II has been provided by the Alfred P. Sloan Foundation, the Participating Institutions, the National Science Foundation, the U.S. Department of Energy, the National Aeronautics and Space Administration, the Japanese Monbukagakusho, the Max Planck Society, and the Higher Education Funding Council for England. The SDSS Web Site is http://www.sdss.org/." }, "0910/0910.3213_arXiv.txt": { "abstract": "Using 11-years of OGLE V-band photometry of Q2237+0305, we measure the transverse velocity of the lens galaxy and the mean mass of its stars. We can do so because, for the first time, we fully include the random motions of the stars in the lens galaxy in the analysis of the light curves. In doing so, we are also able to correctly account for the Earth's parallax motion and the rotation of the lens galaxy, further reducing systematic errors. We measure a lower limit on the transverse speed of the lens galaxy, $v_{\\rm t} > 338\\kms$ (68\\% confidence) and find a preferred direction to the East. The mean stellar mass estimate including a well-defined velocity prior is $0.12 \\leq \\left \\leq 1.94$ at $68\\%$ confidence, with a median of $0.52~M_\\sun$. We also show for the first time that analyzing subsets of a microlensing light curve, in this case the first and second halves of the OGLE V-band light curve, give mutually consistent physical results. ", "introduction": "\\label{sec:intro} Quasar microlensing provides a unique tool for studying the properties of cosmologically distant lens galaxies and the structure of quasars \\citep[see][]{Wambsganss06}. Each of the multiple images of the quasar passes through the gravitational potential of the stars along the line-of-sight in the lens galaxy. These stars microlens each of the ``macro'' images, so the total magnification of each quasar image is strongly affected by the lensing effects of the stars and the size of the quasar emission region. Since the observer, lens galaxy, stars, and source quasar are all moving, these magnifications change on timescales of 1-10 years with order unity amplitudes. The relevant physical scale for quasar microlensing is the Einstein radius projected into the source plane plane, \\begin{eqnarray} \\rE &=& D_{\\rm OS} \\sqrt{\\frac{4G\\left}{c^2}\\frac{D_{\\rm LS}}{D_{\\rm OL}D_{\\rm OS}}} \\nonumber \\\\ &=& 1.8\\times 10^{17} \\left(\\frac{\\left}{M_\\sun}\\right)^{1/2}\\cm, \\end{eqnarray} where $G$ is the gravitational constant, $c$ is the speed of light, $\\left$ is the mean stellar mass of the stars, $D_{\\rm LS}, D_{\\rm OL}$, and $D_{\\rm OS}$ are the angular diameter distances between the lens-source, observer-lens, and observer-source respectively, where we have used the lens and source redshifts for Q2237+0305 \\citep[$z_{\\rm l}=0.0394\\,,z_{\\rm s}=1.685$,][Q2237 hereafter]{Huchra85}. The observed microlensing amplitude is controlled by the ratio between the source size, $R_{\\rm V} \\approx 6\\times10^{15}\\cm$ (in V-band, see our companion paper \\citet{Poindexter09}, hereafter Paper II) and $\\rE$, in the sense that smaller accretion disks produce larger variability amplitudes than larger disks. If a source is much larger than $\\rE$, there is little change in the magnification. The timescale for variability is determined by the relative velocities of the observer, the lens, its stars, and the source. Generally, the lens motions dominate \\citep{Kayser89}, leading to two characteristic timescales. There is the timescale to cross an Einstein radius, \\begin{eqnarray} \\tE &\\approx& \\frac{\\rE}{v_{\\rm lens}} \\frac{(1+z_{\\rm l})\\dol}{\\dos} \\nonumber \\\\ &\\approx& 8 \\left(\\frac{v_{\\rm lens}}{462\\kms}\\right)^{-1} \\left(\\frac{\\rE}{2\\times10^{17}\\cm}\\right) {\\rm yr}, \\nonumber \\\\ \\label{eqn:te} \\end{eqnarray} and there is the timescale to cross the source, \\begin{eqnarray} \\ts &\\approx& \\frac{R_{\\rm V}}{v_{\\rm lens}} \\frac{(1+z_{\\rm l})\\dol}{\\dos} \\nonumber \\\\ &\\approx& 0.4 \\left(\\frac{v_{\\rm lens}}{462\\kms}\\right)^{-1} \\left(\\frac{R_{\\rm V}}{6\\times10^{15}\\cm}\\right) {\\rm yr}, \\nonumber \\\\ \\label{eqn:ts} \\end{eqnarray} where $v_{\\rm lens}$ is the expected transverse speed of the lens for Q2237. These timescales are also affected by the direction of motion relative to the shear \\citep{Wambsganss90}. Variation is guaranteed on timescale $\\tE$ and can be observed on timescale $\\ts$ if the magnification pattern locally has structure on the scale of $R_{\\rm V}$. Quantitative studies using quasar microlensing have exploded in the last few years. Recent efforts have studied the relationships between accretion disk size and black hole mass \\citep{Morgan10}, size and wavelength \\citep{Anguita08,Bate08,Eigenbrod08b,Poindexter08,Floyd09,Mosquera09}, sizes of non-thermal (X-ray) and thermal emission regions \\citep{Pooley07,Morgan08a,Chartas09,Dai09}, and the dark matter fraction of the lens \\citep{Dai09,Pooley09,Mediavilla09}. All these studies have used static magnification patterns which ignore the random stellar motions in the lens galaxy. However, the stellar velocity dispersions of lens galaxies are comparable to the peculiar velocities of galaxies, and ignoring them can lead to biased results. For example, the average direction of motion of the source is the same for all images, but this coherence is limited by the random motions of the stars. With fixed patterns one must either overestimate the coherence, by using the same direction of motion for each image, or underestimate it by using independent directions for each image \\citep[see][]{Kochanek07}. In either case, it would be dangerous to attempt measurements that depend on this coherence, such as disk shape and orientation, or the transverse peculiar velocity of the lens. \\citet{Kundic93}, \\citet{Schramm93}, and \\citet{Wambsganss95} considered the effects of random stellar motions and found that these motions can also lead to shorter microlensing time scales because the pattern velocities of the microlensing caustics can be much higher than any physical velocities. As a result, measurements based on static magnification patterns may underestimate source sizes or mean masses or overestimate the transverse velocity in order to match the effects created by the random stellar motions. \\citet{Wyithe00} showed that it is possible to statistically correct for these effects and that the velocity correction can be up to 40\\% depending on the optical depth and shear. Another benefit to dynamic magnification patterns is the ability to properly account for the velocity of Earth around the Sun and the rotation of the lens galaxy \\citep{Tuntsov04}. Dynamic magnification patterns are also important because they impart a well-measured physical scale to the patterns. All direct microlensing observations are in ``Einstein units'' where the length scale is $\\left^{1/2}\\cm$. Determining masses, velocities, or source sizes requires some sort of dimensional prior. In our studies we have generally followed \\citet{Kochanek04a} and used velocity priors designed to mimic the combined effects of random and ordered motion. The reason for focusing on the velocity is that we know two of the contributions, our velocity and the stellar velocities, and the remaining peculiar velocities of the lens and source are truly random variables for which we have reasonable priors from cosmological models. Source sizes turn out to be little affected by the choice of priors \\citep[see the discussion in][]{Kochanek04a}, but estimates of the true velocity and the mean stellar masses are affected. Hopefully by including the true random stellar motions we can further reduce the sensitivity of microlensing results to such priors. \\begin{figure*} \\plotone{tracks} \\caption{Example of a trial source trajectory (dark line segments) superposed on an instantaneous point-source magnification pattern for $\\mmass = 0.3$. Darker shades indicate higher magnification. An HST H-band image in the center labels the images and the corresponding magnification patterns. Each pattern is rotated to have the correct orientation relative to the lens. This particular LC2 trial has an effective lens-plane velocity of $\\sim 600\\kms$ Northeast. The solid disk at right has a radius of $10^{17}\\cm$ on these patterns.} \\label{fig:tracks} \\end{figure*} In this paper we use microlensing to measure the peculiar velocity of a lens galaxy and the mean mass of its stars including the effects of the stellar motions, Earth's motion, and the rotation of the lens galaxy. The transverse velocity direction can be measured with microlensing because the shear sets a preferred direction for each image and the statistics of variability depend on the motion relative to this axis (see Figure \\ref{fig:tracks}). In theory, accurately measuring the transverse peculiar velocity of many galaxies over a broad range of redshifts could form the basis of a new cosmological test \\citep{Gould95}. Measuring the mean stellar mass, including remnants, is an independent means of checking local accountings \\citep[e.g.][]{Gould00}, which must be assembled from very disparate selection methods for high mass, low mass, evolved and dead, remnant stars. Moreover, doing this is possible in detail only for the Galaxy. While microlensing is relatively insensitive to the mass function \\citep[see][]{pac86,Wyithe00b}, there are some prospects of exploring this in the future as well \\citep[e.g.][]{Wyithe01,Schechter04,Congdon07}. This work expands on the methods described in \\citet{Kochanek04a} and \\citet{Kochanek07} by adding the random stellar motions in the lens galaxy. In this paper we address the computational issues and then apply this improved technique to determine the transverse motion of Q2337 and the mean mass of its stars. In Paper II, we study the shape of the accretion disk of the source quasar. We describe the photometric data in \\S\\ref{sec:data}. Then we describe the Bayesian Monte Carlo Method and the models we use in \\S\\ref{sec:methods}. Our results are presented in \\S\\ref{sec:results} followed by a discussion in \\S\\ref{sec:discussion}. We use an $\\Omega_M = 0.3, \\Omega_{\\Lambda} = 0.7$ flat cosmological model with $H_0 = 72~\\rm km~s^{-1}~Mpc^{-1}$. ", "conclusions": "\\label{sec:discussion} By including the random motions of the stars, we can now use microlensing to study the peculiar velocity of the lens galaxy and to estimate mean stellar masses and potentially the stellar mass functions with fewer systematic uncertainties. In particular, we find a clear preference for the direction of motion of the lens galaxy. In fact, as we use a less restrictive velocity prior, the direction of motion is better constrained since faster speeds are allowed. We cannot however, determine the speed without additional priors. If we assume a mean stellar mass of $\\left = 0.3 M_\\sun$, we find that the peculiar velocity is $v_{\\rm t} < 486 \\kms$ which is consistent with the other estimates \\citep{Wyithe99,GilMerino05} but more fully includes all the physical uncertainties. It should not be surprising that we can determine a dimensionless quantity, the direction of motion, better than the dimensional speed, given the basic problem of microlensing that all observables are in $\\sqrt{\\left}\\cm$. Very roughly (see Figure \\ref{fig:tracks}), the preferred direction has images A and B moving more closely to perpendicular to the ridges of the magnification patterns created by the shear and images C and D are moving more parallel to the shear direction. This is consistent with the variability of A/B compared to C/D. We included the parallax effects of the Earth's orbit, and the results weakly favored its inclusion. We had hoped that modeling the random motions would be more of a help in breaking these degeneracies by setting a physical scale. This is probably true for low mean masses $\\left$. For fixed variability amplitudes, reproducing the light curves with a low mean mass requires small physical velocities, while high mean masses require high velocities. Adding the stellar motions at their observed dispersion eliminates low mass solutions independent of the unknown peculiar velocities by setting a floor to the velocity scale. High mass solutions need peculiar velocities, $\\sigma$, that are larger than the stellar motions, $\\sigma_*$, and so are only constrained by the priors on the peculiar velocities. Essentially, the dynamic patterns act like static patterns once $\\sigma_* < \\sigma$, and we recover the familiar degeneracies of static patterns. Thus, our correct treatment of the stellar motions constrains low mass but not high mass solutions in the absence of a peculiar velocity prior. With a well-defined cosmological prior on $\\sigma$ \\citep{Tinker09}, we find that $0.12 \\leq \\left \\leq 1.94$ at $68\\%$ confidence, demonstrating that the microlensing objects are typical of stellar populations and their remnants. This mass range is consistent with expectations for normal stellar populations (see \\S\\ref{sec:mass}), but not tightly constraining. We largely ignore the macro magnifications predicted by the mass distribution of the lens galaxy in our calculations because of their systematic uncertainties. However some recent studies have made use of this information by analyzing image pairs straddling a critical curve which should have the same magnification \\citep{Floyd09,Bate08}. A concern is that the macro magnification may be affected by undetected substructure, differential extinction, or contamination by the lens or host galaxy. In our standard analysis we use the AC signal and largely discard the DC signal by not tightly constraining the mean magnification. Given sufficiently long light curves, the results will converge to the true magnification offsets. Even for Q2237, with its decade long OGLE light curves, the data are not sampling long enough paths across the patterns (see Figure \\ref{fig:tracks}) to show convergence. At present, the distribution of differential mean magnification offsets are too broad (Figure \\ref{fig:zero}) to tightly constrain any systematic magnification offsets. Fortunately, our results for the other physical parameters are little affected by whether we allow these offsets to vary or constrain them with the extinction estimates of \\citet{Agol09}. Finally, we show for the first time that microlensing variability in a lens gives the same results when analyzing different portions of its light curve. The analysis of light curves LC1 and LC2, corresponding to the 1st and 2nd halves of the 11 year OGLE monitoring period, lead to statistically consistent distributions for every parameter we consider. This both confirms our ability to measure parameters and gives us tighter constraints after combining the results. It would be computationally challenging to analyze the full light curve simultaneously because it becomes (exponentially?) harder to fit longer light curves. However, such full analyses are likely needed for some quantities, particularly the magnification offsets, to converge. In the future we will likely include binaries, even though their effect is not likely to influence the results other than interpreting the meaning of the mean stellar mass (by up to $0.05M_\\sun$, as discussed in \\S\\ref{sec:results}). However, like the projection of our motion relative to the CMB, the streaming velocities in Q2237 are small compared to the peculiar velocities, and so are do little to break the degeneracy. The effects of streaming velocities will be seen most strongly in true disk lenses (none are known, except, potentially PMN~J2004$-$1349, \\citet{Winn01}), or in lenses such as Q~J0158$-$4325 (see \\citet{Morgan08b} for a microlensing analysis of this active system) lying close to the equator of the CMB dipole, which will have the full $369\\kms$ dipole motion \\citep{Hinshaw09}. These CMB equatorial lenses should also show significantly shorter microlensing variability time scales. Detecting this effect would be an independent confirmation of the kinematic origin of the dipole. Q2237 was a natural first candidate for a full analysis with moving stars because of the excellent OGLE data, short microlensing timescales, and negligible time delays between the images. However, there is no problem extending our approach to analyzing microlensing data with moving stars to any other microlensing analysis. Even if the time delays are unknown, cases with different trial delays could simply be tried sequentially \\citep{Morgan08b}. Moreover our method can easily be extended to multi-wavelength data sets to examine the structure of the accretion disk varies with wavelength \\citep{Poindexter08}. The memory requirements would be too great to fit each band simultaneously as in \\citet{Poindexter08}, but we can use a modified version of the method \\citet{Dai09} applied to the joint optical and X-ray analysis of RXJ1131$-$1231. The models are first run on the band with the most epochs. As good fitting trials are found, the starting points, velocities, and $\\chi^2$ matrix are saved. Next, for each successive band, we recompute the light curves corresponding to the epochs and source sizes of the other wavelengths, and the results of these new fits are used to continue the $\\chi^2$ calculation. Since the overall execution times are only modestly longer than using static stars, there is no reason not to use this more physically correct approach." }, "0910/0910.3449_arXiv.txt": { "abstract": "{ The spectrum of cosmic rays (CRs) is affected by their escape from an acceleration site. This may have been observed not only in the gamma-ray spectrum of young supernova remnants (SNRs) such as RX~J1713.7$-$3946, but also in the spectrum of CRs showering on the Earth. } {The escape-limited model of cosmic-ray acceleration is studied in general. We discuss the spectrum of CRs running away from the acceleration site. The model may also constrain the spectral index at the acceleration site and the ansatz with respect to the unknown injection process into the particle acceleration. } { Analytical derivations. We apply our model to CR acceleration in SNRs and in active galactic nuclei (AGN), which are plausible candidates of Galactic and extragalactic CRs, respectively. In particular, for young SNRs, we take account of the shock evolution with cooling of escaping CRs in the Sedov phase. } {The spectrum of escaping CRs generally depends on the physical quantities at the acceleration site, such as the spectral index, the evolution of the maximum energy of CRs and the evolution of {{\\bf the normalization factor of the spectrum. }} It is found that the spectrum of run-away particles can be both softer and harder than that of the acceleration site. } { The model could explain spectral indices of both Galactic and extragalactic CRs produced by SNRs and AGNs, respectively, suggesting the unified picture of CR acceleration. } ", "introduction": "\\label{sec:intro} The origin of cosmic rays (CRs) has been one of the long-standing problems. The number spectrum of nuclear CRs observed at the Earth, $\\mathcal{N} (E) \\propto E^{-s}$, shows a break at the ``knee'' energy ($\\approx10^{15.5}$~eV), below which the spectral index is about $s \\approx -2.7$ \\citep{cronin1999}. Because of the energy-dependent propagation of CRs, the spectral shape at the source is different from that observed at the Earth. Taking into account the propagation effect, the source spectral index has been well constrained as $s \\approx 2.2-2.4$ in various models \\citep[e.g.,][]{strong98,putze2009}. This value of $s$ has been also inferred in order to explain the Galactic diffuse gamma-ray emission \\citep[e.g.,][]{strong00}. This fact may give us valuable insights on the acceleration mechanism of CRs. Mechanisms of CR acceleration have also been studied for a long time and the most plausible process is a diffusive shock acceleration (DSA) \\citep{krymsky77,axford77,bell78,blandford78}. Very high-energy gamma-ray observations have revealed that the existence of high-energy particles at the shock of young supernova remnants (SNRs), which supports the DSA mechanism as well as the paradigm that the Galactic CRs are produced by young SNRs \\citep[e.g.,][]{enomoto02, aharonian04, aharonian05c, katagiri05}. Recent progress of the theory of DSA has revealed that the back-reactions of accelerated CRs are important if a large number of nuclear particles are accelerated \\citep{drury81,malkov01}. There are several observational facts which are consistent with the predictions of such nonlinear model \\citep{vink03, bamba03b, bamba05a, bamba05b, warren2005, uchiyama07, helder2009}. The model predicts, however, the harder spectrum of accelerated particles at the shock than $f (p) \\propto p^{-4}$ (corresponding to $s=2$) where $p$ is the momentum of CRs, in particular, near the knee energy \\citep{malkov97, berezhko99, kang01}. This fact apparently contradicts with the source spectral index of $s \\approx {2.2-2.4}$ inferred from the CR spectrum at the Earth. Even in the test-particle limit of DSA, such a soft source spectrum requires a shock with the small Mach number ($M \\la10$), which is unexpected for young SNRs. There are several models of DSA, depending on the boundary conditions imposed. Different models predict different spectra of CRs dispersed from the shock region. So far, the age-limited acceleration has been frequently considered as a representative case (\\S~\\ref{sec:agelimit}). In this model, all the particles are stored around the shock while accelerated. When the confinement becomes inefficient, all the particles run away from the region at a time. Then, the source spectrum of CRs which has just escaped from the acceleration region is expected to be the same as that at the shock front. Therefore, this model predicts that the source spectrum is the same as that of accelerated particles, which is typically harder than the observed one. In this paper, we consider an alternative model, the escape-limited acceleration, to explain the observed CR spectrum at the Earth (\\S~\\ref{sec:escape}). This model is preferable to the age-limited acceleration when we consider observational results for young SNR RX~J1713.7$-$3946, of which TeV $\\gamma$-ray emission is more precisely measured than any others (\\S~\\ref{sec:rxj1713}). The nature of CRs with energies much higher than the knee energy is also still uncertain. While CRs below the second knee ($\\sim {10}^{17.5}$~eV) may be Galactic origin, the highest energy CRs above $\\sim{10}^{18.5}$~eV are believed to be extragalactic. Possible candidates are active galactic nuclei (AGNs) \\citep[e.g.,][]{BS87,Tak90,RB93,PMM09}, gamma-ray bursts \\citep[][]{Wax95,Vie95,MINN06}, magnetars \\citep[][]{Aro03,MMZ09} and clusters of galaxies \\citep[][]{KRB97,ISMA07}. The intermediate energy range from $\\sim {10}^{17.5}$~eV to $\\sim{10}^{18.5}$~eV is more uncertain. Both the Galactic and extragalactic origins are possible and it may just a transition between the two. As the extragalactic origin, AGNs \\citep{berezinsky06}, clusters of galaxies \\citep{MIN08} and hypernovae \\citep{WRMD07} have been proposed so far. Among these possibilities, AGN is one of the most plausible candidates for accelerators of high-energy CRs, because it can explain the UHECR spectrum above $\\sim 10^{17.5}$~eV assuming the proton composition. In such a proton dip model, the source spectrum of UHECRs is $\\mathcal{N} \\propto E^{-s}$ with $s=2.4$--2.7, depending on models of the source evolution \\citep{berezinsky06}. The required source spectral index of $s=2.4$--2.7 can be explained by several possibilities. First, it can be attributed to the acceleration mechanism itself. One can consider non-Fermi acceleration mechanisms \\citep{berezinsky06} or the two-step diffusive shock acceleration in two different shocks \\citep{Alo+07}. Second, the index can be attributed to a superposition of many AGNs with different maximum energies, and one can suppose that AGNs with different luminosities may have different maximum energies \\citep{KS06}. Recently, \\citet{berezhko2008} proposed another possibility under the cocoon shock model. In this cocoon shock scenario, different maximum energies can be interpreted as maximum energies of escaping particles at different ages of AGN jets. Although it is very uncertain whether efficient CR acceleration occurs there, this scenario would also be one of the possibilities to be investigated in detail. The organization of the paper is as follows. After the brief introduction of the age-limited model of the CR acceleration (\\S~\\ref{sec:agelimit}), we study the escape-limited model in \\S~\\ref{sec:escape}. For a simple understanding, the general argument in a stationary, test-particle approximation is given in \\S~\\ref{app:testparticle}. Then, we derive the formulae of the maximum energy of accelerated particles in \\S~\\ref{sec:pmax_escape}, and of the spectrum of escaping particles in \\S~\\ref{sec:spect_escape}. We consider the applications to young SNR and AGN in \\S~\\ref{sec:snr} and \\S~\\ref{sec:AGN}, respectively. Section~\\ref{sec:discussion} is devoted to a discussion. ", "conclusions": "\\label{sec:discussion} In this paper, we have investigated the escape-limited model of CR acceleration, in which the maximum energy of CRs of an accelerator is limited by the escape from the acceleration site. The typical energy of escaping CRs decreases as the shock decelerates because the diffusion length becomes longer. After revisiting the escape-limited model and reconsider its detail more generally, we have derived a simple relation between the spectrum of escaping particles and one in the accelerator. Then, using the obtained relation, we have discussed which model of injection is potentially suitable to make the Galactic and extragalactic CRs observed at the Earth. For young SNRs, we have considered the shock evolution with cooling by escaping CRs and those spectra for the three injection models. As a result, we have found that in the case PH, it is difficult to satisfy the condition for the source spectrum of Galactic CRs ($s_{\\rm esc} \\approx 2.2$--2.4). On the other hand, $s_{\\rm esc} \\approx 2.2$--2.4 can be achieved in cases of PS and TL. We have also applied our escape-limited model to AGN cocoon shocks as well as young SNRs. This model is just one of the various candidates proposed so far, even if AGNs are UHECR accelerators. Nevertheless, it is interesting that the young SNR and the AGN cocoon shock scenarios can explain the Galactic and extragalactic cosmic rays observed at the Earth in the same picture for all the three injection models if we accept the proton-dip model inferring $s_{\\rm esc} \\approx2.4$--2.7. Whether the proton dip model is real or not can also be tested by future UHECR and high-energy neutrino observations. In this paper, we have focused on the proton case. Obviously, heavier nuclei become important above the knee so that we need to take into account them in order to explain the CR spectrum over the whole energy range. We can also apply the escape-limited model to heavy nuclei CRs for this purpose, although it is beyond the scope of this paper. We point out a potential problem for the magnetic field amplification in the escape-limited model. In the case of young SNRs, we have determined the evolution of the maximum energy in the phenomenological way, and adopted $\\alpha \\approx 6.5$. Using Eqs.~(\\ref{eq:pmax_esc2}) and (\\ref{eq:pm_general}), we obtain $B \\propto u_{\\rm sh}^{\\frac{10+2c}{3}}$, where $\\ell \\propto R_{\\rm sh}^c$. The same result is obtained for AGN cocoon shocks because we have considered the case in which the same parameters describe both the young SNR shocks and the AGN cocoon shocks. In particular, for $c =1$ as is used in Eq. (31), we obtain $B^2 \\propto u_{\\rm sh}^8$ which means that $B$ rapidly decreases with radius (or time). In principle, both the dependence of $B$ on $u_{\\rm sh}$ and the value of $c$ can be determined theoretically, and then the evolution of the maximum energy should be predicted. Some previous works are based on theoretical arguments on the magnetic field evolution \\citep[e.g.,][]{bell04,PLM06,berezhko2008,caprioli09}, which seem to be different from our phenomenological one. At present, the mechanisms of the particle acceleration and the magnetic field amplification are still highly uncertain despite of many theoretical efforts \\citep[e.g.,][]{niemiec08,requelme09, ohira09a, LM09}. Hence, we expect that further theoretical and observational studies might reveal this discrepancy or exclude the possibility of escape-limited acceleration in the future. In this paper, we have mainly considered spectra of dispersed CRs around young SNRs and AGN cocoon shocks. However, applications to other astrophysical objects are, of course, possible. For example, the old SNRs detected by {\\it Fermi} LAT, such as W28, W44, W51 and IC443 \\citep{abdo09,abdo09w51c} have been of great interest because they likely generate escaping CRs. In fact, the number of CRs around such old SNRs is likely to decrease with time or the shock radius, that is $\\beta<0$ while $\\alpha>0$. For example, when we consider the dynamics of an old SNR, we have $u_{\\rm sh} \\propto t^{-2/3} \\propto R_{\\rm sh}^{-2}$ \\citep[e.g.,][]{yamazaki06}, so that we have $E \\propto u_{\\rm s}^2R_{\\rm s}^3\\propto R_{\\rm sh}^{-1}$, i.e., $\\delta = -1 < 0$. On the other hand, the value of $\\alpha$ may be different from 6.5, which could be attributed to various complications such as the interaction with the dense molecular cloud, and so on. For example, for the maximum hardening case, that is, $s_\\mathrm{esc}=s + (s-1)(\\sigma^{-1}-1)/\\alpha$ (see Appendix A.2), we find $s_{\\rm esc} \\approx 1.5$ when $\\alpha \\approx 1$ and $s \\approx 2.5$ where we assume $s=(\\sigma +2)/(\\sigma -1)$. This might be the case for old SNRs such as W51C \\citep{abdo09w51c}. In addition, the maximum energy may be rather small for the old SNRs, so that the spectrum above $p_m^{(\\rm esc)}$ would be suppressed. The spectrum of high-energy gamma rays might give us important information on both the acceleration and escape processes of CRs with energies much lower than the knee energy." }, "0910/0910.2353_arXiv.txt": { "abstract": "We present the results of simulations of shadows cast by a zone plate telescope which may have one to four pairs of zone plates. From the shadows we reconstruct the images under various circumstances. We discuss physical basis of the resolution of the telescope and demonstrate this by our simulations. We allow the source to be at a finite distance (diverging beam) as well as at an infinite distance (parallel beam) and show that the resolution is worsened when the source is nearby. By reconstructing the zone plates in a way that both the zone plates subtend the same solid angles at the source, we obtain back high resolution even for sources at a finite distance. We present simulated results for the observation of the galactic center and show that the sources of varying intensities may be reconstructed with accuracy. Results of these simulations would be of immense use in interpreting the X-ray images from recently launched CORONAS-PHOTON satellite. ", "introduction": "\\label{intro} Zone Plate Telescopes (ZPTs) have generated immense theoretical interest in the past. (Ables, 1968; Dicke, 1968; Barrett \\& Swindell, 1996; Desai, Norris \\& Nemiroff, 1993; Desai et al. 1993; 1998, 2000). Due to the availability of technology to make finer zones using X-ray and gamma-ray opaque materials, the interest of using zone plates have increased recently. In Chakrabarti et al. (2009, hereafter Paper I), we presented extensive theoretical studies and some of the results of the experiments conducted at the Indian Centre for Space Physics X-ray laboratory on zone plate telescopes. Such telescopes have been used in Indian payload system RT-2 aboard the Russian satellite CORONAS-PHOTON which was launched on 30th January, 2009. As mentioned in Paper I the ZPTs have an advantage over other high resolution X-ray telescopes in that they can have arbitrarily high angular resolution and that the resolution can also be independent of energy bands in a large range of energy. The only disadvantage is of course that the ZPT is a two element system as opposed to the conventional coded aperture masks (CAMs) which are single element systems. In the present paper of our series we shall present the results of the Monte-Carlo simulations of the nature of photon distribution on the detector planes for various combinations of zone plates and X-ray sources. The basic starting point in this paper would be Paper I and references therein. Here, we consider one pair to four pairs of zone plate combinations, with a point source and distributed sources. We also vary the distance of source from finite to infinite. We hope to be able to present the most comprehensive results in this subject so that interpretation of the data from CORONAS-PHOTON may become easier. In our next paper (Nandi et al. 2009) the instrumentation and early results from RT-2 payload which uses ZPTs would be presented. The plan of the present paper is the following: In the next Section, we discuss the angular resolution of a ZPT and also the variation of the nature of the reconstructed image when the source distance varies. In Section 3, we will present in detail the results of our simulations in various circumstances. Finally, in Section 4, we make concluding remarks. ", "conclusions": "In this Paper, we presented results of various numerical simulations obtained by varying the source, the zone plate telescope parameters and the detector parameters. We considered both two-pair and four-pair configurations which enabled us to remove the pseudo-source and both the pseudo-source and DC-offset respectively. We pointed out that the pseudo-source cannot be totally removed if the source itself is at a finite distance. This is due to the fact that at a finite distance, different pair is hit by photons from the source at various angles and the contributions to the pseudo-source do not cancel out. We showed that the resolution of the instrument is worsened when the source is placed at a finite distance. However, we showed that if the telescope is modified in a way that both the plates subtend equal angles at the source, the high resolution is recovered. Of course, a practical difficulty of such a method is that the second zone plate has to be modified dynamically as the source distance is varied, though, for instance for medical purposes, one could keep the source at a fixed distance always and thus a fixed sized ZP2 would suffice. Alternatively, one could convolve the distorted fringes obtained due to sources at a finite distance by theoretical Moir\\'{e} pattern obtained for the source at infinite distance before deconvolving the pixel information. This is beyond the scope of the present paper and will be dealt with elsewhere. We studied the cases with both CMOS and CZT detectors and showed that the setup with CZT detectors will have very limited field of view in order to even reconstruct a single source. This is because of the large pixel size. With a CMOS detector we simulated how a zone plate telescope would view the center of the Galaxy. When there are multiple sources of varying intensity, we showed that one could first obtain a general image and then improve upon it by subtracting the fringes from the strongest ones. The zone plate telescopes have very high potential for future space astronomy, especially since it will have higher resolutions over a large range of energy. They may also be used for medical science since the sources at finite distances can also be resolved well if modified ZPTs are used. Recently ZPTs have been used in RT-2/CZT payload aboard Russian satellite CORONAS-PHOTON for the first time. The satellite has been recently launched and the results are expected in near future, especially when the sun becomes active. The results of the zone plate imagers will be discussed elsewhere (Nandi et al. 2009)." }, "0910/0910.5059_arXiv.txt": { "abstract": "{} {Recently a new analysis of cluster observations in the Milky Way found evidence that clustered star formation may work under tight constraints with respect to cluster size and density, implying the presence of just two sequences of young massive cluster. These two types of clusters each expand at different rates with cluster age.} {Here we investigate whether similar sequences exist in other nearby galaxies.} {We find that while for the extragalactic young stellar clusters the overall trend in the cluster-density scaling is quite comparable to the relation obtained for Galactic clusters, there are also possible difference. For the LMC and SMC clusters the densities are below the Galactic data points and/or the core radii are smaller than those of data points with comparable density. For M83 and the Antenna clusters the core radii are possibly comparable to the Galactic clusters but it is not clear whether they exhibit similar expansion speeds. These findings should serve as an incentive to perform more systematic observations and analysis to answer the question of a possible similarity between young galactic and extragalactic star clusters sequences.} {} ", "introduction": "\\begin{figure}[t] \\resizebox{\\hsize}{!}{\\includegraphics[angle=-90]{Pfalzner_fig1.ps}} \\caption{{Cluster density as a function of cluster size for clusters more massive than 10$^3$ \\Msun as published by Pfalzner (2009). references therein.}} \\label{fig:cluster_rad} \\end{figure} Lada \\& Lada (2003) showed that in our Galaxy most stars do not form in isolation but in cluster environments. These young clusters typically consist of thousand stars or more with densities ranging from $<$0.01 to several 10$^5$ \\Msun pc$^{-3}$ The wide variety of observed cluster densities lead to the assumption that clusters are formed over this entire density range. Albeit Maiz-Apellaniz (2001) noticed the existance of two types of clusters in the Galaxy, only recently Pfalzner (2009) found that massive clusters develop in a bimodal way as in Fig.\\ref{fig:cluster_rad}, where the red, green and blue symbols represent clusters with ages $t_c <$ 4 Myr, 4 Myr $ < t_c <$ 10 Myr, 10 Myr $< t_c <$ 20 Myr, respectively. Two well-defined sequences in the density-radius plane emerge showing the {\\em bi-modal nature of the cluster evolution.} \\begin{figure}[t] \\resizebox{\\hsize}{!}{\\includegraphics[angle=0]{LMC_size_density.eps}} \\caption{a) Cluster density as a function of cluster radius and b) cluster size as a function of cluster age. Here the LMC and SMC (purple circle) clusters are shown as filled symbols and the Milky Way as open symbols (same clusters as in Fig.~\\ref{fig:cluster_rad}). The values for the clusters in the LMC are taken from Mackey \\& Gilmore 2003a and McLaughlin \\& van der Marel 2005, the SMC value from Mackey \\& Gilmore 2003b.} \\label{fig:LMC_cluster_size_age} \\end{figure} Pfalzner (2009) classified the two modes as starburst and leaky cluster sequences. The starburst cluster sequence implies a population of compact clusters (0.1 pc) with high initial densities (10$^5$ - 10$^6$ \\Msun pc$^{-3}$) which then expand with the mass-density decreasing as $\\sim R^{-3}$ (an actual $2\\sigma$-fit gives a $R^{\\alpha}$-dependence with $\\alpha$=-2.71 $ \\pm $ 0.32) and evolve to have sizes of a few pc over a period of 10 Myr or longer. Prominent members of this type of cluster are for example Arches, NGC 3603 and Westerlund 1. The leaky cluster sequence implies the creation of a second population of diffuse clusters ($\\sim$ 5pc) which expand loosing mass during the process, until they have sizes of a few tens of pc. The expansion also follows a predefined track, but here the density decreases as $\\sim R^{-4}$ (actual fit value $\\alpha=-$4.07$\\pm0.33$). Here NGC 6611, Ori 1a-c or U Sco are typical examples. Note, the cluster radii were not determined exactly the same way for all clusters shown in Fig.~\\ref{fig:cluster_rad}. However, star burst cluster values all represent the core radii apart from $\\chi$ Per and h Per (for a discussion on the determination of these two radii see Pfalzner 2009). It follows that star formation occurs only under an extremely limited set of conditions, and may require a fundamental revision of star formation hypotheses in our Galaxy. This immediately raises the questions of the origin of these two distinct cluster sequences and whether similar density-radius correlations could be found in other galaxies. This Letter addresses the latter question. Most extragalactic clusters one observes are likely to belong to the starburst cluster sequence because although their extension is on average smaller the luminosity of their high number of O-stars is much easier to detect than the smaller number of O stars spreading over a larger area in leaky clusters. So in all figures subsequent to Fig.~1 only the starburst clusters of the Galaxy are shown for comparison and the radial extent is the core radius. \\begin{table*} \\begin{center} \\begin{tabular}{l|*{5}{l}} \\hline \\\\[-2ex] Galaxy & radius determination & mass determination & background & age determination\\\\\\hline \\\\[-2ex] LMC$^1$ & King profile fit & from mass/light ratio from & mean surface brightness & combination of IMF $^8$ \\\\ & half-surface brightness & evolutionary code$^5$ & in annulus at r $\\gg R_c$ & and evolutionary code$^5$ \\\\ SMC$^2$ & King profile fit & from mass/light ratio from & mean surface brightness & Based on color-magnitude \\\\ & half-surface brightness & evolutionary code$^5$ & in annulus at r $\\gg R_c$ & diagram$^{12}$ \\\\ Antennae$^3$ & Fit with ISHAPE code$^6$ & UBVI-colors and mass/light & Fit with ISHAPE code & LICK line strength indices \\\\ & & ratio$^7$ with extinction- & & and $^{11}$ \\\\ & & corrected SSP & & \\\\ M 83$^4$ & Fit with ISHAPE code$^6$ & Isochromes$^9$ & Fit with ISHAPE code & UVBI-colors, 2-color and \\\\ & & & & S-sequnce diagram $^{10}$ \\end{tabular} \\caption{Determination methods of radius, mass and age and background consideration ($^1$Maxkey \\& Gilmore 2003a, $^2$ Mackey \\& Gilmore 2003b, $^3$ Mengel et al. 2008, $^4$ Larsen \\& Richter 2004, $^5$Fioc \\& Rocca-Volmerangen 1997, $^6$ Larsen 1999, $^7$Anders \\& Fritz-von-Alvensleben 2003, $^8$ Kroupa et al. 1993, $^9$ Girardi 2000, $^{10}$ Girardi 1995, $^11$ Schweizer 2004 $^{12}$ Chiosi 1995. \\label{table:method}} \\end{center} \\end{table*} Currently the Milky Way is not in a intense star formation phase accounting for the scarcity of starburst clusters in the Galaxy. Another reason is that extinction at low latitudes hampers the detection of distant Galactic clusters. By contrast, there are starburst and interacting galaxies such as the \"Antennae\", where many clusters with masses $>$ 10$^4$ \\Msun\\ are observed with ages $<$1Gyr (Zhang \\& Fall 1999). In between, there exist the dwarf starburst galaxies such as NGC 4214, where several massive clusters with ages $<$100 Myr are visible in the central regions. This Letter shows a first investigation of whether similar development tracks can be found for clusters in such different galaxies. Observationally one is restricted to nearby galaxies, and even there radii are only resolved for a very limited sample. Here we scanned the literature for clusters younger than 30 Myr in the Large Magellanic Cloud (LMC), Small Magellanic Cloud (SMC), Messier 83 (M83) and the Antennae. ", "conclusions": "For the extragalactic young stellar clusters the overall trend in the cluster radius versus cluster density relation seems to fit the relation obtained for Galactic starburst clusters by Pfalzner (2009). A possible difference is indicated for the LMC clusters, where the densities are mostly below the comparable Galactic data at equal radius, or the radii are smaller than those of data points with comparable density. This might indicates that the sizes for LMC clusters are systematically smaller than the Galactic cluster sizes. This could be explained by intrinsic observational effects. However, the sample size of extragalactic young clusters ($<$ 20 Myr ) with known radii and masses is so small that this investigation can only be regarded as a first hint that the young starburst cluster sequence might also exist in galaxies other than the Milky way. In this study only literature values of age, mass and radius were used mostly obtained by different methods. Further studies with a larger sample size and more homogenous data acquisition will be required to confirm this result." }, "0910/0910.5868_arXiv.txt": { "abstract": "\\noindent Timing noise in the data on accretion-powered millisecond pulsars (AMP) appears as irregular pulse phase jumps on timescales from hours to weeks. A large systematic phase drift is also observed in the first discovered AMP \\sax. To study the origin of these timing features, we use here the data of the well studied 2002 outburst of \\sax. We develop first a model for pulse profile formation accounting for the screening of the antipodal emitting spot by the accretion disk. We demonstrate that the variations of the visibility of the antipodal spot associated with the receding accretion disk cause a systematic shift in Fourier phases, observed together with the changes in the pulse form. We show that a strong secondary maximum can be observed only in a narrow intervals of inner disk radii, which explains the very short appearance of the double-peaked profiles in \\sax. By directly fitting the pulse profile shapes with our model, we find that the main parameters of the emitting spot such as its mean latitude and longitude as well as the emissivity pattern change irregularly causing small shifts in pulse phase, and the strong profile variations are caused by the increasing inner disk radius. We finally notice that significant variations in the pulse profiles in the 2002 and 2008 outbursts of \\sax\\ happen at fluxes differing by a factor of 2, which can be explained if the inner disk radius is not a simple function of the accretion rate, but depends on the previous history. ", "introduction": "The first accreting millisecond X-ray pulsar (hereafter AMP) \\sax\\ was discovered in 1998 \\citep{WvdK98, CM98}. Since then, 12 objects of this class have been discovered, with the spin frequencies in the range 182--599 Hz. Many AMPs (e.g., IGR J00291+5934, XTE J1751$-$205, XTE J1807$-$294, \\sax, XTE J1814$-$338) often show nearly sinusoidal profiles, consistent with only one visible hotspot, probably because the accretion disk extends very close to the stellar surface. The general stability of the pulse profile allows to obtain a high photon statistics and to use the average pulse profile to get constraints on the neutron star mass-radius relation and the equation of state \\citep{PG03}. However, sometimes the profiles do show variations. In \\sax, for example, a dip appears around the pulse maximum at high photon fluxes, and at low fluxes the profile becomes skewed to the left and sometimes is double-peaked. The pulse profile evolution during its five observed outbursts is amazingly similar, with significant changes in the profiles accompanied by large (0.2 cycles) jumps in the phase of the fundamental (see \\citealt{BDM06,HPC08,HPC09,PW09}; \\citealt{IP09}, hereafter IP09). Besides these noticeable profile changes, the timing noise with timescales from hours to weeks is known to be present in the data of this and other AMPs, when the measured pulse phase delays show fluctuations around a mean trend \\citep{pap07,CCHY08,RDB08,PWvdK09}. There are many possible reasons for the pulse profile changes and the timing noise: changes in the spot shape and position, variations of the emission pattern, inner disk radius or the optical depth of the accretion stream \\citep{p08}. Changes in the position of the hotspots \\citep{lamb08,PWvdK09} can cause random phase irregularities; however, it is difficult to imagine that they can lead to qualitative changes in the pulse form (such as transformation of a single-peaked profile to a double-peaked) unless accompanied by strong variation in the emissivity pattern. On the other hand, the visibility of the secondary spot might change even with small fluctuations of the accretion rate. IP09 proposed that the strong pulse profile variations in the 2002 outburst of \\sax\\ results from the appearance of the secondary antipodal spot to our view when the magnetospheric radius increases and the accretion disk recedes from the neutron star with the dropping accretion rate. The decreasing amplitude of Compton reflection in the spectra during the outburst supports this interpretation. In this Letter, we directly fit the pulse profiles of \\sax\\ during its 2002 outburst with the theoretical model for AMP pulses including the effect of partial screening of the antipodal emitting spot by the accretion disk. We aim at quantifying the variations in the position of the hotspot centroid, its emissivity pattern and the inner accretion disk radius, and determining the causes of the variations in the pulse shape and the timing noise. \\begin{figure} \\centerline{\\epsfig{file= f1.eps,width=8cm}} \\caption{Pulse profiles produced by two antipodal spots of various shapes for $\\rin$ varying from 20 to 30 km with step 2.5 km (from bottom to top). The left hand side panels correspond to the blackbody emissivity pattern $h=0$, and the panels at the right are for $h=0.8$. At lowest radii, the view of antipodal spot is almost fully blocked by the disk, while at highest radii, the whole spot is visible. Stellar parameters: $M=1.4M_\\odot$, $R=10.3$ km, $i=65\\degr$, $\\theta=10\\degr$. The spot angular size $\\rho$ is given by equation (\\ref{eq:rhorin}), and for the ring and crescent geometries, the inner radius is $\\rho/\\sqrt{2}$. } \\label{f:profile} \\end{figure} ", "conclusions": "We have developed a model for the pulse profile formation in AMPs which accounts for the partial screening of the antipodal emitting spot by the accretion disk. We have demonstrated that the appearance of the antipodal spot, due to the increasing inner disk radius, leads to sharp changes in the pulse profile and corresponding jump in Fourier phases. We showed that a strong secondary maximum appears only in a rather narrow intervals of inner disk radii, explaining the very short appearance of the double-peaked profiles in \\sax. By directly fitting the model to the pulse shapes of \\sax\\ observed during its 2002 outburst, we were able to quantify the variations in the position of the spot centroids, their emissivity pattern and the inner accretion disk radius. The sharp jumps in Fourier phases, observed together with the strong changes in the pulse form, can be explained by the variations of the visibility of the antipodal spot associated with the receding accretion disk. We also note that a factor of 2 difference in fluxes in 2002 and 2008 outbursts of \\sax, when the pulse profile started to change significantly, implies that the inner disk radius is not a simple function of the accretion rate, but might depend on the previous history. We finally note that many physical timing noise mechanisms probably operate in AMPs simultaneously." }, "0910/0910.0160_arXiv.txt": { "abstract": "s{I know better than to come between the experts here assembled and their research programs, so I confine these remarks to lessons to be drawn on the state of our subject from the histories of research in three Windows on the Universe: cosmology, our extragalactic neighborhood, and life in other worlds.} ", "introduction": " ", "conclusions": "" }, "0910/0910.5603_arXiv.txt": { "abstract": "{ The \\emph{INTEGRAL} satellite is discovering a large population of new X-ray sources which were missed by previous missions due to high obscuration and, in some cases, very short duty cycles. The nature of these sources must be addressed via the characterization of their optical and/or infrared counterparts. } {We investigate the nature of the optical counterparts to five of these newly discovered X-ray sources.} {We combine infrared spectra in the $I$, $J,H$ and $K$ bands together with $JHK$ photometry to characterize the spectral type, luminosity class and distance to the infrared counterparts to these systems. For IGR~J19140$+$0951, we present spectroscopy from the red to the $K$ band and new red and infrared photometry. For SAX~J18186$-$1703 and IGR~J18483$-$0311, we present the first intermediate-resolution spectroscopy reported. Finally, for IGR~J18027$-$2016, we present new $I$ and $K$ band spectra.} {We find that four systems harbour early-type B supergiants. All of them are heavily obscured, with $E(B-V)$ ranging between 3 and 5, implying visual extinctions of $\\sim$ 9 to 15 magnitudes. We refine the published classifications of IGR~J18027$-$2016 and IGR~J19140$+$0951 by constraining their luminosity class. In the first case, we confirm the supergiant nature and rule out class III. In the second case, we propose a slightly higher luminosity class (Ia instead of Iab) and give an improved value of the distance based on new optical photometry. Two other systems, SAX~J18186$-$1703 and IGR~J18483$-$0311 are classified as Supergiant Fast X-ray Transients (SFXTs). XTE~J1901$+$014, on the other hand, contains no bright infrared source in its error circle.} {Owing to their infrared and X-ray characteristics, IGR~J18027$-$2016 and IGR~J19140$+$0951, emerge as Supergiant X-ray binaries with X-ray luminosities of the order of $L_{X}\\sim [1-2]\\times 10^{36}$ erg s$^{-1}$, while SAX~J1818.6$-$1703 and IGR~J18483$-$0311, turn out to be SFXTs at 2 and 3 kpc, respectively. Finally, XTE~J1901$+$014 emerges as a puzzling source: its X-ray behaviour is strongly reminiscent of the SFXTs but a supergiant nature is firmly ruled out for the counterpart. We discuss several alternative scenarios to explain its behaviour. } ", "introduction": "The {\\it INTEGRAL} satellite has broadened our knowledge of X-ray binaries, by discovering a significant population of new X-ray sources. The Third {\\it IBIS/ISGRI} cataloge \\citep{bird07} contains more than 420 sources of which 167 are new. The on-line cataloge of {\\it INTEGRAL} sources\\footnote{\\tiny{\\texttt{http://isdcul3.unige.ch/$\\sim$rodrigue/html/igrsources.html}}} lists 245 sources discovered or re-discovered by {\\it INTEGRAL}. Of these, 61 ($\\sim 25$ \\%) remain unidentified and 81 ($\\sim 33$ \\%) are extragalactic objects (AGNs, Seyferts, QSOs, etc). The rest are mainly X-ray binaries. 51 of them ($\\sim 21$ \\%) turn out to be High Mass X-ray Binary candidates (HMXB), which were missed by previous missions because of their high obscuration, because of their very short X-ray cycles or a combination of both. Approximately half of them have been confirmed so far. Amongst these new 51 HMXB systems, there are a few Be/X-ray binaries (BeXB), Supergiant X-ray binaries (SGXB) and the newly established class of Supergiant Fast X-ray Transients \\citep[SFXT; e.g.,][]{smith06,neg06a,sguera06}. This last class comprises 16 candidates of which nearly half have been confirmed. Before we can conclusively classify every new source and understand its X-ray behavior, we must first establish the nature of the counterpart. Since the majority of these sources are located in the galactic plane, and concentrated towards the galactic center region, they suffer from large extinction ($A_{V}\\lesssim 15$ mag). Thus, their detection in the visual is often very difficult, and other strategies must be adopted. Detection and study of the counterparts in the infrared, however, is possible with 4~m class telescopes. An important caveat, though, is that the spectral classification of hot stars using infrared spectra is much more uncertain than using the well established MK standard 3950--4750 \\AA\\ region. For example, the spectral classification based on the $K$ band spectrum cannot be done without fundamental ambiguities because of the lack of adequate spectral features in that range \\citep{hanson96}. This problem can be, at least partially, circumvented by using the combined information of several spectral bands. However, this is often impossible to achieve within a single observing campaign. Following the discovery by \\emph{INTEGRAL}, intense observing campaigns have been undertaken \\citep[amongst others]{chaty08,masetti08,nespoli08,neg09}. As these discoveries proceed, it is fundamental to refine and establish the nature of these new sources. In particular, a very important parameter is the distance to the system which, in turn, gives the X-ray luminosity, the main observable used to constrain the theoretical models. Most of the work done, however, has been based on $H$ and/or $K$ band spectra to overcome the high extinction. Classification based on a single infrared spectrum (in many cases of intermediate or low resolution only) allows only a crude approximation. For example, the difference in $M_{V}$ between a B1\\,Ib and a B1\\,Ia supergiant (which are difficult or even impossible to tell apart based on an intermediate resolution $H$ or $K$ spectrum) results in an uncertainty of an order of magnitude error in $L_{X}$ \\citep{neg09}. In this paper, we combine new spectroscopy in several bands with new $JHK$ photometry (also $RI$ for one source) to further refine the spectral classification of a sample of these sources. ", "conclusions": "\\begin{table*} \\caption[]{Derived parameters for stars in our sample, calculated using the observed data. } \\label{tab:data} \\begin{center} \\begin{tabular}{llccccccrr} \\hline \\hline \\noalign{\\smallskip} Source & Spectral type & Class & $E(B-V)$ & $E(J-K)$ & $M_{V}$ & $M_{K}$ & $d$ (kpc) & $L_{X}$ (erg/s)$^{\\rm (a)}$ \\\\ \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} IGR~J18027$-$2016 & B1\\,Ib & SGXB & 3.04$\\pm0.02$ & 1.54 & $-5.8$ & $-5.1$ & 12.4$\\pm 0.1$ & (1.65$\\pm0.03$) $\\times 10^{36}$ \\\\ SAX~J18186$-$1703 & B0.5\\,Iab & SFXT & 5.08$\\pm0.05$ & 2.54 & $-6.4$ & $-5.65$ & 2.1$\\pm0.1$ & (8$\\pm0.4$) $\\times 10^{35}$ \\\\ IGR~J18483$-$0311 & B0-1\\,Iab & SFXT & 5.22$\\pm0.02$& 2.61 & $-6.4$ & $-5.66$ & 2.83$\\pm 0.05$ & (1.9$\\pm0.06$) $\\times 10^{36}$ \\\\ IGR~J19140$+$0951 & B0.5\\,Iab-a & SGXB & 5.5$\\pm0.1$ & 2.75 & -6.65$\\pm 0.03$ & -5.90$\\pm0.25$ & 3.6$\\pm 0.04$ & (1.6$\\pm0.3$) $\\times 10^{36}$ \\\\ \\hline \\noalign{\\smallskip} XTE~1901$+$014 & G5\\,III - A3\\,V & LMXB? & $\\sim$3.7 & 1.8 & & & 2-7 & 1$^{+1}_{-0.8}\\times 10^{38}$ \\\\ \\noalign{\\smallskip} \\hline \\end{tabular} \\begin{list}{}{} \\item[$^{\\mathrm{a}}$] Observed peak luminositites for the energy ranges (2--10)\\,keV, (22--55)\\,keV, (20--100)\\,keV, (2--20)\\,keV and (3--12) \\,keV respectively. \\end{list} \\end{center} \\end{table*} Four of our sources have early-B supergiant companions. All four counterparts turn out to cover a rather narrow spectral interval, being early-B (B0--B1.5) supergiants of moderate or high luminosity. This identifies the nature of the X-ray source unambiguously as HMXBs. Furthermore, they are all heavily obscured, with $E(B-V)\\sim 3-5.5$, implying extinctions on the order of $A_{V}=9-16$ mag in the visual band. But, on the other hand, XTE~1901+014 has no obvious counterpart in any band except $K^{\\prime}$. Thus, a supergiant nature is definitely excluded, thereby setting it apart from the other four. IGR~J18027$-$2017 is a persistent X-ray source with a counterpart around B1\\,Ib. It is therefore, a SGXB system. The reason why it has not been detected by earlier X-ray missions is not clear. For example, it was not detected by \\emph{ROSAT}. At a distance of $d\\approx 12.4$ kpc, its {\\it XMM-Newton} detected flux in the 2--10~keV band for the main pulse of $8.9\\times 10^{-11}\\:{\\rm erg}\\,{\\rm s}^{-1}\\,{\\rm cm}^{-2}$, would yield a luminosity of $L_{X}=1.65\\times 10^{36}\\:{\\rm erg}\\,{\\rm s}^{-1}$. This is typical for these kind of systems albeit in the lower end. Since the detected flux outside the primary pulse is a little bit smaller, the system can spend a fraction of its time in the upper $\\times 10^{35}$ erg s$^{-1}$. Given its high obscuration, this can explain why it was missed by earlier surveys. For SAX~J18186$-$1703, the NIR counterpart is a B0.5Iab star, confirming its nature as a SFXT. The 22--50\\,keV flux reported by \\citet{zurita09} is of the order of (2--8)$\\times 10^{-11}\\:{\\rm erg}\\,{\\rm s}^{-1}\\,{\\rm cms}^{-2}$ in quiescence while it reaches (1--2)$\\times 10^{-9}\\:{\\rm erg}\\,{\\rm s}^{-1}\\,{\\rm cms}^{-2}$ at the strongest flares. At the deduced distance of $\\sim 2.1$ kpc, the X-ray luminosity of this object would be $\\sim 3\\times 10^{34}\\:{\\rm erg}\\,{\\rm s}^{-1}$ in quiescence, while the peak luminosity can be as high as $\\sim 8\\times 10^{35}\\:{\\rm erg}\\,{\\rm s}^{-1}$. This is lower than those found in other SFXTs ($\\sim 10^{36}\\:{\\rm erg}\\,{\\rm s}^{-1}$). \\citet{zurita09} have found that this system is in a very eccentric orbit ($e\\sim 0.3-0.4$), with $P_{orb}\\approx 30$ d, amongst the largest yet found for SFXT, reaching a periastron distance between 2 and 3 $R_{*}$. This would locate the compact object at a slightly larger distance from the donor than in the SGXB systems ($a\\lesssim 2R_{*}$) thereby reducing the $L_{X}$ of the outbursts slightly below $10^{36}\\:{\\rm erg}\\,{\\rm s}^{-1}$ as is observed. IGR~J18483$-$0311 is a transient system with a B0--B1\\,Iab primary. It has been classified as an intermediate SFXT \\citep{rahoui08}. Like many other SFXTs, it is a nearby source at $\\sim$ 2.8 kpc. For this distance, the {\\it INTEGRAL}/IBIS 20--100~keV flux \\citep{sguera07} translates into a luminosity $\\sim 1.9 \\times 10^{36}\\:{\\rm erg}\\,{\\rm s}^{-1}$, for the strongest flares, which is typical for this kind of objects. The orbital period of this system is 18.5 d, while the spectral type of the companion is similar to that of SAX~J18186$-$1703. Therefore, the compact object would be located slighly closer to the star and, correspondingly, its average X-ray luminosity will be higher, in agreement with the scenario described in \\citet{neg08}. IGR~J19140$+$0951 is a persistent X-ray source and therefore, owing to its counterpart, a SGXB located at $d\\approx 3-4$~kpc, a little bit further than previously thought. For the bright {\\it INTEGRAL} flux in the 2-2-0~keV band of $1\\times 10^{-9}\\:{\\rm erg}\\,{\\rm s}^{-1}\\,{\\rm cms}^{-2}$ \\citep{hannika04}, the $L_{X}=1-2\\times 10^{36}\\:{\\rm erg}\\,{\\rm s}^{-1}$, typical of a SGXB albeit, again, in the lower end. The fact that, together with IGR 18027-2017, they have lower X-ray fluxes than the average SGXBs explains their more recent detections in the newer, and deeper {\\it INTEGRAL} observations. As can be seen in Fig.~\\ref{fig:galaxy}, the sources tend to concentrate in or behind the section of the Sagittarius arm projected onto the direction of the Galactic Bulge, an area intensively scanned by {\\it INTEGRAL}. They are, thus, relatively close to the Sun (in comparison to typical distances to HMXBs), although highly reddened. This fact can introduce a bias in the newly detected populations of obscured sources, which tend to harbor SG companions, as only the brighter and/or closer ones can be reached with telescopes of moderate apertures in the 4~m class. A transient system with a main sequence companion, like a BeXB, will be too faint to be conclusively classified \\citep[i.e.][]{zurita08}. Remarkably, the furthest source in our sample is the least reddened one. This raises the question of the true abundance of HMXBs (and OB stars in fact) in our Galaxy. As can be seen in Table \\ref{tab:data}, the high reddening is not a sufficient condition to be an SFXT since some SGXBs (like IGR~J19140$+$0951) are more reddened. On the other hand, it is not a \\emph {necessary} condition either, as other SFXTs present rather low reddenings \\citep[for example IGR~J08408$-$4503 with $E(B-V)=0.5$][]{neg09}. Furthermore, the position in the arms of the Galaxy of the newly discovered HMXBs (including the SFXTs) is entirely consistent with the positions of the previously known HMXBs (including SGXBs). Therefore there is no reason to believe that they form a different population. This would be expected from a scenario like that discussed in \\citet{neg08}. In this scenario, the SFXTs are explained as just SGXBs where the compact object is located further out in the system, in a region where the stellar wind of the supergiant companion has become substantially clumped. The number of HMXBs in the Galaxy would then be much larger than previously thought. But so would the number of isolated OB stars whose much lower X-ray emission will not show up in the current X-ray surveys if they are heavily obscured. This is supported by the discovery, in the last few years, of a growing population of very massive infrared clusters \\citep[e.g.][]{crow06, mess08}. Therefore, in principle, there is no reason to believe that the relative abundance of HMXBs, with respect to OB stars, differs from the current values which are well explained by the theoretical models. Finally, XTE~J1901$+$014 emerges as a transient source whose X-ray behaviour is characteristic of the SFXTs but where the presence of a Supergiant companion is clearly ruled out. It can still be a massive star close to the main sequence in the outskirts of the Galaxy, perhaps with a $16\\,M_{\\sun}$ BH as a companion. In this respect, it is interesting to note the case of the LMC transient A\\,0535$-$66 reaching $L_{X}\\approx 8\\times 10^{38}\\:{\\rm erg}\\,{\\rm s}^{-1}$ \\citep[e.g.,][]{pp81}. A similar source at the edge of the Galaxy could have a similar hard X-ray behaviour. However, taking into account the moderate hardness (the source is seen as 1RXH~J190140.1+012630) and obscuration of the source, it seems more likely that the counterpart is a late type star located between 2~kpc and the Long Galactic bar (7~kpc), somewhere between luminosity class V to class III, but definitely not consistent with a class I star. Whether a main sequence or giant late type star wind can produce the instabilities necessary to produce the observed outbursts via accretion onto a compact object remains an issue." }, "0910/0910.2812_arXiv.txt": { "abstract": "% CCD photometric observations of the Algol-type eclipsing binary X Tri have been obtained. The light curves are analyzed with the Wilson-Devinney (WD) code and new geometric and photometric elements are derived. A new O-C analysis of the system, based on the most reliable timings of minima found in the literature, is presented and apparent period changes are discussed with respect to possible and multiple Light-Time Effect (LITE) in the system. Moreover, the results for the existence of additional bodies around the eclipsing pair, derived from the period study, are compared with those for extra luminosity, derived from the light curve analysis. ", "introduction": "The system was observed during 5 nights from October 2008 to January 2009 at the Athens University Observatory, using a 40-cm Cassegrain telescope equipped with the CCD camera ST-8XMEI and B, V, R, I Bessell filters. The light curves (hereafter LCs) were analysed with the $PHOEBE~ 0.29d$ software \\citep{PZ05} which uses the 2003 version of the WD code \\citep{WD71,W79}. Due to the lack of spectroscopic mass ratio of the system, the q-search method was applied in Mode 2, 4 and 5 in order to find the most probable value of the (photometric) mass ratio (q). The least squares method with statistical weights has been used for the analysis of the O-C diagram. The current O-C diagram of X Tri includes 572 times of minima taken from the literature. The following ephemeris: $Min.I=HJD~2422722.285 + 0.9715341^{d} \\times E$ \\citep{K01} was used as the initial one for the O-C analysis of the compiled times of minima. ", "conclusions": "The results of the LCs solution (see figure 1 and table 1) show that X Tri is a semi-detached system with the secondary star filling its Roche Lobe. The periodic variations of the orbital period of the system could be explained by adopting the existence of three additional components, which were found to have minimal masses 0.18, 0.24 and 0.22, respectively (see figure 1 and table 2). An extra light contribution to the luminosity of the EB was taken into account in the LCs solution but it was found to be less than 1\\%. Such a small extra luminosity could be explained by the small values of the minimal masses of the possible additional components found. The steady decrease rate of its period is probably due to angular momentum loss, since the direction of the flow (from the more massive to the less massive component), revealed from the O-C diagram analysis, comes in disagreement with the one derived from the LCs analysis. \\begin{table} \\caption{The parameters of X Tri derived from the LCs solutions} \\label{tab1} \\scalebox{0.58}{ \\begin{tabular}{ccccccc} \\hline \\textbf{Parameter} & \\textbf{value} & \\textbf{Parameter} & \\multicolumn{4}{c}{\\textbf{value}} \\\\ \\hline $q~(m_{2}/m_{1})$ & 0.599~(2) & & B & V & R & I \\\\ $i$~(deg) & 87.9~(1) & $x_{1}$$^{***}$ & 0.551 & 0.478 & 0.402 & 0.322 \\\\ $T_1$$^{**}$(K), $T_2$ (K) & 8600, 5188~(4) & $x_{2}$$^{***}$ & 0.835 & 0.692 & 0.597 & 0.503 \\\\ $A_1$$^*$, $A_2$$^*$ & 1, 0.5 & $L_{1}/L_{T}$ &0.893~(2)&0.839~(2)&0.795~(2)&0.739~(1) \\\\ $g_1$$^*$, $g_2$$^*$ & 1, 0.32 & $L_{2}/L_{T}$ &0.107~(1)&0.160~(2)&0.201~(2)&0.246~(3) \\\\ $\\Omega_{1}$, $\\Omega_{2}$ & 4.27~(1), 3.06 & $L_{3}/L_{T}$ &0.000~(1)&0.000~(1)&0.004~(1)&0.015~(2) \\\\ $\\chi^{2}$ & 1.278 & & \\\\ \\hline \\textit{$^*$assumed},&\\textit{$^{**}$ \\citet{G83},} &\\textit{$^{***}$\\citet{VH93}},& \\textit{$L_{T} = L_{1}+L_{2}+L_{3}$} \\end{tabular}} \\end{table} \\begin{table} \\caption{The results of the O-C diagram analysis for X Tri} \\label{tab2} \\scalebox{0.52}{ \\begin{tabular}{cccccc} \\hline \\textbf{Parameters of the EB} & \\textbf{value} & \\textbf{Parameters of the LITEs} & \\multicolumn{3}{c}{\\textbf{value}} \\\\ \\hline $M_1^{*}+M_2~(M_\\odot)$ & 2.1 + 1.26 & & $3^{rd} body$ & $4^{th} body$ & $5^{th} body$ \\\\ $Min. I$~(HJD) & 2442502.731~(2) & $P$~(yrs) & 36.9~(5) & 22.4~(3) & 16.8~(4) \\\\ $P$~(days) & 0.9715318~(2) & $T_0$~(HJD),~$\\omega$~(deg) &2452916~(373), 220~(98)&2455069~(335), 34~(13)& --, -- \\\\ $c_{2}~(\\times 10^{-10})$ & -2.0308~(2) & $A$~(days),~$e$ & 0.0052~(3), 0.2~(2) & 0.0040~(4), 0.5~(1) & 0.003~(2), 0.0 \\\\ $\\dot{P}~(\\times 10^{-10})$ & -1.5269~(2) & $M_{min}~(M_\\odot)$ & 0.18~(1) & 0.24~(1) & 0.22~(1) \\\\ \\hline \\textit{$^*$assumed} \\end{tabular}} \\end{table} \\begin{figure}[ht!] \\plottwo{LC.eps}{O-C.eps} \\caption{Left panel: The synthetic LCs (black solid lines) along with the observed ones (colored points). Right panel: The multiperiodic fitting (red solid line) on the O-C points (black points) (upper part) and the O-C residuals (lower part).} \\end{figure}" }, "0910/0910.4025_arXiv.txt": { "abstract": "{The majority of stars that leave the main sequence are undergoing extensive mass loss, in particular during the asymptotic giant branch (AGB) phase of evolution. Observations show that the rate at which this phenomenon develops differs highly from source to source, so that the time-integrated mass loss as a function of the initial conditions (mass, metallicity, etc.) and of the stage of evolution is presently not well understood. } {We are investigating the mass loss history of AGB stars by observing the molecular and atomic emissions of their circumstellar envelopes.} {In this work we have selected two stars that are on the thermally pulsing phase of the AGB (TP-AGB) and for which high quality data in the CO rotation lines and in the atomic hydrogen line at 21 cm could be obtained.} {V1942 Sgr, a carbon star of the Irregular variability type, shows a complex CO line profile that may originate from a long-lived wind at a rate of $\\sim$ 10$^{-7}$ \\Msold, and from a young (\\lsim10$^4$\\,years) fast outflow at a rate of $\\sim$ 5 10$^{-7}$ \\Msold. Intense \\HI emission indicates a detached shell with 0.044 \\Msol ~of hydrogen. This shell probably results from the slowing-down, by surrounding matter, of the same long-lived wind observed in CO that has been active during $\\sim$\\,6\\,10$^{5}$ years. On the other hand, the carbon Mira V CrB is presently undergoing mass loss at a rate of 2\\,10$^{-7}$\\,\\Msold, but was not detected in H\\,{\\sc {i}}. The wind is mostly molecular, and was active for at most 3\\,10$^{4}$ years, with an integrated mass loss of at most 6.5\\,10$^{-3}$\\,\\Msol.} {Although both sources are carbon stars on the TP-AGB, they appear to develop mass loss under very different conditions, and a high rate of mass loss may not imply a high integrated mass loss.} ", "introduction": "Low- to intermediate-mass stars, at the end of their main-sequence evolution, become first hydrogen shell-burning red giants (RGB $-$Red Giant Branch$-$ stars), then hydrogen and helium shell-burning red giants (AGB $-$Asymptotic Giant Branch$-$ stars). In this second phase they may undergo mass loss at a very large rate ($>$ 10$^{-8}$ \\Msold), even so large that it has a decisive effect on their late evolution (Olofsson 1999). They are surrounded by expanding envelopes of gas and dust that have been extensively observed with radio molecular lines and infrared continuum emission. These tracers are used to estimate mass-loss rates. However the estimates are somewhat ambiguous because the mass-loss rate of a given source may vary on many different timescales. The mass change as a function of time due to mass loss is thus difficult to evaluate, and to connect with stellar evolution models. Furthermore molecular lines probe an extent of the circumstellar shell (CS) that is limited by photo-dissociation, and therefore furnish information mainly on the inner parts of CSs, and on the recent mass loss. To try to circumvent these difficulties we have started a systematic programme of observations of red giants in the line of atomic hydrogen at 21 cm. We have published some of our results in several recent papers, and first reports on sizeable samples have been presented by G\\'erard \\& Le~Bertre (2006, hereafter GL2006) and Matthews \\& Reid (2007, hereafter MR2007). A major difficulty of this programme is the confusion caused by the 21 cm emission from the Interstellar Medium (ISM) that is located on the same line-of-sight as the source of interest. This has a strong impact on the observations which have to be conducted with a specific approach, and on the data processing that aims at providing spectra corrected for the ISM emission. Perhaps more confounding, as circumstellar matter is expected to be at some stage injected in the ISM, the confusion by the {\\ub {local}} ISM might actually be at least partly of stellar origin, i.e. caused by material ejected at an earlier stage of evolution. In addition to observing systematically the \\HI line at 21 cm in a large sample of sources with different properties, it is also important to choose objects for which the Galactic confusion is low and/or can be tracked easily, and therefore corrected accurately. The detailed study of such spectra should serve as a guide to exploit the data that are obtained in more difficult situations. Here we present our results on two carbon stars, V1942 Sgr and V CrB, for which the confusion is not a serious problem, and which have \\HI properties that differ radically. Both are N-type carbon stars (CGCS 4229 and CCCS 2293, respectively) and have already been detected in CO rotational lines (Olofsson et al. 1993a). However the only available CO spectrum of V1942 Sgr had a poor signal-to-noise ratio, and for our study it appeared essential to also obtain new CO data of high quality. ", "conclusions": "The combination of high velocity resolution CO and \\HI data is a promising tool to investigate the history of mass loss by evolved stars. The low level of Galactic \\HI emission and the absence of small-scale structure in this emission have allowed us to obtain \\HI data of high quality on V1942 Sgr and V CrB with the NRT. We have also obtained high quality CO (1-0) and (2-1) spectra of V1942 Sgr with the IRAM 30-m telescope. For V1942 Sgr, the CO spectra exhibit composite profiles, that reveal a low velocity wind of $\\sim$ 10$^{-7}$ \\Msold ~and a high velocity wind, possibly bipolar, of $\\sim$ 5\\,10$^{-7}$ \\Msold. The comparison with the \\HI spectrum shows that this high velocity wind is recent with an age of at most 10$^4$ years. On the other hand, the low velocity wind appears to have filled a cavity of $\\sim$ 0.2 pc in radius and built the detached shell, that was discovered by IRAS, over a period of 5 10$^5$ years. Follow-up observations with the VLA and ALMA would help to improve this scenario, or possibly lead to a new scheme. The narrowness of the \\HI line profile in V1942\\,Sgr brings new evidence that AGB stellar winds are slowed down by their surrounding medium as surmised by Young et al. (1993b). For V CrB, the CO spectra that have been published reveal an outflow with expansion velocity, 6.5 \\kms, and mass loss rate, 2.1 10$^{-7}$ \\Msold. The non-detection in \\HI of V CrB sets an upper limit of 3.2 10$^4$ years for the age of this outflow. In such case of a star with low effective temperature, molecular hydrogen data are obviously needed to constrain better the history of mass loss." }, "0910/0910.1436_arXiv.txt": { "abstract": "Due to its unique long-term coverage and high photometric precision, observations from the Kepler asteroseismic investigation will provide us with the possibility to sound stellar cycles in a number of solar-type stars with asteroseismology. By comparing these measurements with conventional ground-based chromospheric activity measurements we might be able to increase our understanding of the relation between the chromospheric changes and the changes in the eigenmodes. In parallel with the Kepler observations we have therefore started a programme at the Nordic Optical Telescope to observe and monitor chromospheric activity in the stars that are most likely to be selected for observations for the whole satellite mission. The ground-based observations presented here can be used both to guide the selection of the special Kepler targets and as the first step in a monitoring programme for stellar cycles. Also, the chromospheric activity measurements obtained from the ground-based observations can be compared with stellar parameters such as ages and rotation in order to improve stellar evolution models. ", "introduction": "During recent years there has been a growing interest in understanding stellar cycles in general and the solar cycle in particular. The reason for this is partly the indications that the solar activity cycle could cause climate change (see e.g. Svensmark \\& Friis-Christensen 1997) and partly that solar cycle 24 now seems to be around two years delayed compared to theoretical predictions (Zimmerman 2009). The ``Sounding stellar cycles with Kepler'' programme will monitor stellar cycles from the amplitude and frequency shifts of the oscillation modes observed in these stars. By comparing these measurements with conventional ground-based chromospheric activity measurements we might be able to increase our understanding of the relation between the chromospheric changes and the changes in the eigenmodes. Asteroseismic analysis of the Kepler observations also allows us to test the hypothesis put forward by B{\\\"o}hm-Vitense (2007) -- that stellar cycles in active stars are generated by the gradient in the rotation rate close to the surface, whereas the stellar cycles in inactive stars are generated by the gradient in the rotation rate at the base of the convection zone. This can be done as the asteroseismic analysis of the Kepler observations will allow us to measure the depth of the convection zone and give estimates of the internal rotation profile. The details of the ``Sounding stellar cycles with Kepler'' programme are described by Karoff et al. (2009). The ground-based chromospheric activity measurements can be used not only to compare the changes on the stellar surfaces to those in the interior over the stellar cycles and to test stellar dynamo models, but also to calibrate rotation-activity-age relations. This can be done as asteroseismology offers a unique possibility to measure ages of solar-type field stars to a precision of 5--10\\% (Christensen-Dalsgaard et al. 2007) and because asteroseismology provides the possibility to measure the inclination of the stellar rotation axis so that the spectroscopic measured $v$sin$i$ can be converted into equatorial surface rotation rates (Gizon \\& Solanki 2003). Also, the first ground-based chromospheric activity measurements that we present here on the potential candidates for observations for the whole mission can be used to rate the different candidates in order to ensure that both active and inactive stars are observed. \\begin{figure}[t] \\begin{center} \\includegraphics[width=3.4in]{karoff.fig01.eps} \\caption{Two spectra illustrating stars with and without chromospheric emission. The wavelength region covers the region around the Ca~{\\sc ii} K line. The upper star (SAO 48081) clearly shows emission in the middle of the Ca~{\\sc ii} K absorption line which proves that this star contains a high level of chromospheric activity. The lower star (SAO 31656) shows no emission which indicates that this star is inactive. The relative intensities have been normalized to the mean intensity in the echelle order and the spectrum for SAO 48081 has been shifted one unit upwards. No radial velocities have been calculated for the stars, but the spectra have been arbitrary aligned.} \\label{fig1} \\end{center} \\end{figure} ", "conclusions": "" }, "0910/0910.3605_arXiv.txt": { "abstract": "The SXI telescope is one of the three instruments on board EXIST, a multiwavelength observatory in charge of performing a global survey of the sky in hard X-rays searching for Supermassive Black Holes. One of the primary objectives of EXIST is also to study with unprecedented sensitivity the most unknown high energy sources in the Universe, like high redshift GRBs, which will be pointed promptly by the Spacecraft by autonomous trigger based on hard X-ray localization on board. The recent addition of a soft X-ray telescope to the EXIST payload complement, with an effective area of ~950 cm$^2$ in the energy band 0.2-3 keV and extended response up to 10 keV will allow to make broadband studies from 0.1 to 600 keV. In particular, investigations of the spectra components and states of AGNs and monitoring of variability of sources, study of the prompt and afterglow emission of GRBs since the early phases, which will help to constrain the emission models and finally, help the identification of sources in the EXIST hard X-ray survey and the characterization of the transient events detected. SXI will also perform surveys: a scanning survey with sky coverage $\\sim2$~$\\pi$ and limiting flux of $\\sim5\\times10^{-14}$ cgs plus other serendipitous. We give an overview of the SXI scientific performance and also describe the status of its design emphasizing how it has been derived by the scientific requirements. ", "introduction": "\\label{sect:sections} The potential of surveys in high energy astronomy is being fully demonstrated by the sky observations performed at X- and soft $\\Gamma$-ray wavelengths during the last decade. In particular XMM Newton$^1$ with its deep sensitivity in the soft X-ray band, INTEGRAL$^2$ and Swift$^3$ up to $>100$ keV have started to reveal the demographics of sources in the near Universe by studying populations of objects. Still lacking today are both a comprehensive picture of transient phenomena (other than GRBs) spanning a broad range of durations and a deep study of the supermassive objects and their host galaxies against Universe age. Studies of transients require fast repointing and improved sensitivity for proper investigations of their prompt emission. The fast follow-up capability of Swift allows to probe the farthest regions of the Universe by the study of Gamma-Ray Bursts (GRB) and also to discover new populations of transients in our Galaxy. GRBs are today the only beacons that can be used to observe the Universe at $z>3-4$, as demonstrated by Swift e.g. with the latest detection of a GRB at z=8. Swift has opened a new window on the far Universe with GRBs, but many more objects are needed including AGNs. Swift also carries an X-ray telescope which is an unvaluable tool to observe the GRB afterglows (and Swift has demonstrated that almost all the GRBs have X-ray afterglows). The Swift/XRT telescope$^4$ is also a powerful tool for localization of the sources seen by Swift/BAT and INTEGRAL/IBIS$^5$ at hard X-ray energies. In the case of IBIS, of the 167 new sources in the 3rd IBIS catalog$^6$ 129 have been identified through optical and NIR spectroscopy. Of these, more than 60\\% have been localized by XRT with typical accuracy $<2-3$ arcsec. Most of the reasons for this success is the possibility for XRT to perform many, relatively short observations. The concept of operability of Swift is also adopted in the design of the Energetic X-ray Imaging Survey Telescope (EXIST) mission$^7$. EXIST is a proposed hard X-ray imaging all-sky deep survey mission and was recommended by 2001 Report of the Decadal Survey. It is a strong candidate to be the Black Hole Finder Probe, one of the three \"Einstein Probes\" in the Beyond Einstein Program, now proposed for the Astro2010 Decadal Survey. EXIST will be launched in a Low Earth Orbit (LEO) and its primary instrument is the High Energy Telescope (HET), a wide field coded aperture instrument covering the 5-600 keV energy band and imaging sources in a $\\sim70\\times90$deg$^2$ field of view with 2~arcmin resolution and better than 20~arcsec positions$^8$. The energy band of HET overlaps with the soft X-ray range covered by the proposed Soft X-ray Imager (SXI), 0.1-10 keV with an effective area of 950cm$^2$ at 1.5 keV and 3.5m focal length. At longer wavelengths, the IRT$^9$ is an optical-IR aperture telescope covering the $0.3-2.2$ micron range with variable spectral resolution and high sensitivity (AB=24 in 100s). The IRT pixel size is 0.15~arcsec and its Field-of-View in Imaging mode is 16~arcmin$^2$. So EXIST is a real multiwavelength observatory for observations of GRBs and Supermassive Black Holes (SMXB) as well as of many other types of transients and high energy sources. In the following we will address some of the topics that will be contributed by the SXI telescope in the study of the high energy sources seen by HET and its capabilities for sky survey at soft X-ray wavelengths. Section 2 contains a brief description of the SXI telescope and its effective area, Section 3 and 4 will describe some of the science cases relevant to SXI. \\begin{figure} \\begin{center} \\begin{tabular}{c} \\includegraphics[height=11cm]{Immagine1_a.eps} \\end{tabular} \\end{center} \\caption[example] { \\label{fig:sxi_telescope} Design overview of the SXI telescope. The instrument (top) has a mirror focal length of 3.5m and the overall length is 4.5m, 70 cm max width. The primary mirror system and the camera (bottom) are also shown. } \\end{figure} ", "conclusions": "\\label{sect:sections} The SXI telescope on board EXIST will take full advantage of the operational strategy adopted for the mission, mostly based on surveys and fast follow-up of GRB and transients. With SXI, EXIST is a real multiwavelength observatory with a sensitive broadband coverage at high energies: 0.1-600 keV, with SXI providing an effective area of 950cm$^2$ at 1.5keV. SXI will perform wide area surveys (scanning \\& serendipitous) and sensitive observation of transient events in the X-ray (e.g. GRB afterglows, AGN flares). It will also help the identification of HET sources detected during the survey, study the absorbed (even Compton thick) AGN Universe and zoom on AGN states. The heritage of the Swift/XRT, XMM-Newton and INTEGRAL allows to conclude that the SXI performance is appropriate to the profile of the EXIST mission as currently designed. Main improvements to Swift/XRT are: factor $\\sim10$ effective area, fast detector readout allowing operation during scanning survey." }, "0910/0910.4251.txt": { "abstract": "{The growth processes from protoplanetary dust to planetesimals are not fully understood. Laboratory experiments and theoretical models have shown that collisions among the dust aggregates can lead to sticking, bouncing, and fragmentation. However, no systematic study on the collisional outcome of protoplanetary dust has been performed so far so that a physical model of the dust evolution in protoplanetary disks is still missing.} {We intend to map the parameter space for the collisional interaction of arbitrarily porous dust aggregates. This parameter space encompasses the dust-aggregate masses, their porosities and the collision velocity. With such a complete mapping of the collisional outcomes of protoplanetary dust aggregates, it will be possible to follow the collisional evolution of dust in a protoplanetary disk environment.} {We use literature data, perform own laboratory experiments, and apply simple physical models to get a complete picture of the collisional interaction of protoplanetary dust aggregates.} {In our study, we found four different kinds of sticking, two kinds of bouncing, and three kinds of fragmentation as possible outcomes in collisions among protoplanetary dust aggregates. Our best collision model distinguishes between porous and compact dust. We also differentiate between collisions among similar-sized and different-sized bodies. All in all, eight combinations of porosity and mass ratio can be discerned. For each of these cases, we present a complete collision model for dust-aggregate masses between $10^{-12}$ and $10^2$~g and collision velocities in the range $10^{-4} \\ldots 10^4~\\rm cm~s^{-1}$ for arbitrary porosities. This model comprises the collisional outcome, the mass(es) of the resulting aggregate(s) and their porosities.} {We present the first complete collision model for protoplanetary dust. This collision model can be used for the determination of the dust-growth rate in protoplanetary disks.} ", "introduction": "} The first stage of protoplanetary growth has still not been fully understood. Although our empirical knowledge on the collisional properties of dust aggregates has considerably widened over the past years \\citep{BlumWurm:2008}, there is no self-consistent model for the growth of macroscopic dust aggregates in protoplanetary disks (PPDs). A reason for such a lack of understanding is the complexity in the collisional physics of dust aggregates. Earlier assumptions of perfect sticking have been experimentally proven false for most of the size and velocity ranges under consideration. Recent work also showed that fragmentation and porosity play important roles in mutual collisions between protoplanetary dust aggregates. In their review paper, \\citet{BlumWurm:2008} show the complex diversity that is inherent to the collisional interaction of dust aggregates consisting of micrometer-sized (silicate) particles. This complexity is the reason why the outcome of the collisional evolution in PPDs is still unclear and why no `grand' theory on the formation of planetesimals, based on firm physical principles, has so far been developed. The theoretical understanding of the physics of dust aggregate collisions has seen major progress in recent decades. The behavior of aggregate collisions at low collisional energies -- where the aggregates show a fractal nature -- is theoretically described by molecular dynamics simulations of \\citet{DominikTielens:1997}. The predictions of this model -- concerning aggregate sticking, compaction, and catastrophic disruption -- could be quantitatively confirmed by laboratory collision experiments of \\citet{BlumWurm:2000}. Also, the collision behavior of macroscopic dust aggregates was successfully modeled by a smooth particle hydrodynamics method, calibrated by laboratory experiments \\citep{GuettlerEtal:2009a, GeretshauserEtal:preprint}. These simulations were able to reproduce bouncing collisions, which were observed in many laboratory experiments \\citep{BlumWurm:2008}. However, as laboratory experiments have shown, collisions between dust aggregates at intermediate energies and sizes are characterized by a plethora of outcomes: ranging from (partial) sticking, bouncing, mass transfer, to catastrophic fragmentation \\citep[see][]{BlumWurm:2008}. From this complexity, it is clear that the construction of a simple theoretical model that agrees with all these observational constraints is very challenging. However, in order to understand the formation of planetesimals, it is imperative to describe the entire phase-space of interest, i.e., to consider a wide range of aggregate masses, aggregate porosities, and collision velocities. Likewise, the collisional outcome is a key ingredient of any model that computes the time evolution of the dust size distribution. These collisional outcomes are mainly determined by the collision velocities of the dust aggregates, and these depend on the disk model, i.e. the gas and material density in the disk and the degree of turbulence. Thus, the choice of the disk model (including its evolution) is another major ingredient for dust evolution models. These concerns lay behind the approach we adopt in this and subsequent papers. That is, instead of first `funneling' the experimental results through a (perhaps ill-conceived) theoretical collision model and then to calculate the collisional evolution, we will directly use the experimental results as input for the collisional evolution model. The drawback of such an approach is of course that experiments on dust aggregate collisions do not cover the whole parameter space and therefore need to be extrapolated by orders of magnitude, based on simple physical models which accuracy might be challenged. However, we feel that this drawback is more than justified by the prospects that our new approach will provide: through a direct mapping of the laboratory experiments, collisional evolution models can increase enormously in their level of realism. In Paper I, we will classify all existing dust-aggregate collision experiments for silicate dust, including three additional original experiments not published before, according to the above parameters (Sect. \\ref{sec:exp-review}). We will show that we have to distinguish between nine different kinds of collisional outcomes, which we physically describe in Sect. \\ref{sec:exp_types}. For the later use in a growth model, we will sort these into a mass-velocity parameter space and find that we have to distinguish between eight regimes of porous and compact dust-aggregate projectiles and targets. We will present our collision model in Sect. \\ref{sec:collision_regimes} and the consequences for the porosities of the dust aggregates in Sect. \\ref{sec:porosities}. In Sect. \\ref{sec:conclusion}, we conclude our work and give a critical review on our model and the involved necessary simplifications and extrapolations. In Paper II \\citep{ZsomEtal:2009}, we will then, based upon the results presented here, follow the dust evolution using a recently invented Monte-Carlo approach \\citep{ZsomDullemond:2008} for three different disk models. This is the first fully self-consistent growth simulation for PPDs. The results presented in Paper II represent the state-of-the-art modeling and will give us important insight into questions, such as if the meter-size barrier can be overcome and what the maximum dust-aggregate size in PPDs is, i.e. whether pebbles, boulders, or planetesimals can be formed. ", "conclusions": "} In the previous sections we have developed a comprehensive model for the collisional interaction between protoplanetary dust aggregates. The culmination of this effort is Fig. \\ref{fig:colored_regimes}, which presents a general collision model based on 19 different dust-collision experiments, which will be adopted in Paper II. Since it plays a vital role, it is worth a critical appraisal. In a few examples, we want to discuss the main simplifications and shortcomings of our current model. (1) The categorization into collisions between similar-sized and different-sized dust aggregates (see Figs. \\ref{fig:categorization} and \\ref{fig:colored_regimes}) is well-motivated as we pointed out in Sect. \\ref{sec:collision_regimes}. However, we may ask ourselves whether this binarization is fundamentally correct, if we need more than two categories, or `soft' transitions between the regimes. At this stage, a more complex treatment would be impractical due to the lack of experiments treating this problem. (2) The binary treatment of porosity (i.e. $\\phi < \\phi_\\mathrm{c}$ for `porous' and $\\phi \\geq \\phi_\\mathrm{c}$ for `compact' dust aggregates) is also a questionable assumption. Although we see fundamental differences in the collision behavior when we use, e.g., porous or compact targets, there might be a smooth transition from the more `porous' to the more `compact' collisions. In addition to that, the assumed value $\\phi_\\mathrm{c} = 0.4$ is reasonable but not empirically affirmed. On top of that, the maximum compaction that a dust aggregate can achieve in a collision depends on many parameters, such as, e.g., the size distribution of the monomer grains \\citep{BlumEtal:2006} and the ability of the granular material to creep sideways inside a dust aggregate \\citep{GuettlerEtal:2009a}. (3) Although the total number of experiments, upon which our model is based, is unsurpassedly large, the total coverage of parameter space (see the experiment boxes in Fig. \\ref{fig:colored_regimes}) is still small. Thus, we sometimes apply extrapolations into extremely remote parameter-space regions. Although not quantifiable, it must be clear that the error of each extrapolation grows with the distance to the experimentally confirmed domains (i.e. the boxes in Fig. \\ref{fig:colored_regimes}). Clearly, more experiments are required to fill the parameter space, and the identification of the key regions in the mass-velocity plane is exactly one of the goals of Paper II. (4) With such new experiments, performed at the `hot spots' predicted in Paper II, we will not only close gaps in our knowledge of the collision physics of dust aggregates but will most certainly reveal completely new effects. The rather simple \\cc\\ panel in Fig. \\ref{fig:colored_regimes} as compared to the more complex \\pC\\ is due to the fact that there are hardly any experiments that back-up the \\cc\\ regime, whereas in the \\pC\\ case we have a rather good experimental coverage of the parameter space. In summary, the sophisticated nature of our collision model is both its strength and its weakness. The drawbacks of identifying four parameters that shape the collision outcome are that rather crude approximations and extrapolations have to be made. However, to acknowledge the role of, e.g., porosity through a binary treatment is still better than to not treat this parameter at all. Our new collision model represents the first attempt to include all existing laboratory experiments (for the material properties of interest); collisional evolution models can enormously profit from this effort. \\subsection{The Bottleneck for Protoplanetary Dust Growth\\label{sec:outlook}} In this paper, we have presented the framework and physical background for an extended growth simulation. What is to be expected from this? Here, we can speculate under which conditions growth in PPDs is most favorable. A view on Fig. \\ref{fig:colored_regimes} immediately shows that large dust aggregates can preferentially grow for realistic collision velocities in the \\cC\\ and \\pC\\ collision regimes (and to a lesser extent in the \\pc\\ case), due to \\Sd. For this to happen, a broad mass distribution of protoplanetary dust must be present. This prerequisite for efficient growth towards planetesimal sizes has also been suggested by \\citet[][see their Fig. 11]{TeiserWurm:2009a}. Agglomeration experiments with micrometer-sized dust grains and a sticking probability of unity (experiments 1 -- 3 in Table \\ref{tab:experiments}) have shown that nature chooses a rather narrow size distribution for the initial fractal growth phase. If this changes when the physical conditions leave no room for growth under quasi-monodisperse conditions, i.e. whether nature is so `adaptive' and `target-oriented' to find out that growth can only proceed with a wide size distribution, will be the subject of Paper II, in which we apply the findings of this paper to a collisional evolution model. \\subsection{Influence of the Adopted Material Properties\\label{sec:material_influence}} The choice of material in our model is 1.5~$\\mu$m diameter silica dust as most of the underlying experiments were performed with this material. Many experiments \\citep{BlumWurm:2000, LangkowskiEtal:2008, BlumWurm:2008} showed that this material is at least in a qualitative sense representative for other silicatic materials -- also for irregular grains with a broader size distribution. Still, the grain size of the dust material may have a quantitative influence on the collisional outcomes. For example, dust aggregates consisting of 0.1~$\\mu$m are assumed to be stickier and more rigid \\citep{WadaEtal:2007, WadaEtal:2008, WadaEtal:2009}, because the grain size may scale the rolling force or breaking energy entering into Eqs. \\ref{eq:S1_threshold} and \\ref{eq:S2_threshold}. However, due to a lack of experiments with smaller monomer sizes, we cannot give a scaling for our model for smaller monomer sizes at this point. Moreover, organic or icy material in the outer regions of PPDs or oxides and sintered material in the inner regions may have a big impact on the collisional outcome, i.e. in enhancing the stickiness of the material and thereby potentially opening new growth channels. As for organic materials, \\citet{KouchiEtal:2002} found an enhanced sticking of cm-sized bodies covered with a 1~mm thick layer of organic material at velocities as high as 500~\\cms\\ and a temperature of $\\sim250$~K. Also icy materials are likely believed to have an enhanced sticking efficiency compared to silicatic materials. \\citet{HatzesEtal:1991} collided 5~cm diameter solid ice spheres, which were covered with a 10 -- 100~$\\mu$m thick layer of frost. They found sticking for a velocity of 0.03~\\cms, which is in a regime where our model for refractory silicatic material predicts bouncing (see \\pp\\ or \\cc\\ in Fig. \\ref{fig:colored_regimes}). Sintering of porous dust aggregate may occur in the inner regions near the central star or -- triggered by transient heating events \\citep[e.g. lightning,][]{GuettlerEtal:2008} -- even further out. Ongoing studies with sintered dust aggregates \\citep{Poppe:2003} show an increased material strength (e.g. tensile strength) by an order of magnitude (C. G\\\"uttler \\& J. Blum, unpublished data). This would at least make the material robust against fragmentation processes and qualitatively shift them from the porous to the compact regime in our model -- without necessarily being compact. Due to a severe lack on experimental data for all these materials, it is necessary and justified to restrict our model to silicates at around 1~AU while it is to be kept in mind that these examples of rather unknown materials might potentially favor growth in other regions in PPDs." }, "0910/0910.1093_arXiv.txt": { "abstract": "We provide fits to the distribution of galaxy luminosity, size, velocity dispersion and stellar mass as a function of concentration index $C_r$ and morphological type in the SDSS. (Our size estimate, a simple analog of the SDSS {\\tt cmodel} magnitude, is new: it is computed using a combination of seeing-corrected quantities in the SDSS database, and is in substantially better agreement with results from more detailed bulge/disk decompositions.) We also quantify how estimates of the fraction of `early' or `late' type galaxies depend on whether the samples were cut in color, concentration or light profile shape, and compare with similar estimates based on morphology. Our fits show that ellipticals account for about 20\\% of the $r$-band luminosity density, $\\rho_{L_r}$, and 25\\% of the stellar mass density, $\\rho_*$; including S0s and Sas increases these numbers to 33\\% and 40\\%, and 50\\% and 60\\%, respectively. The values of $\\rho_{L_r}$ and $\\rho_*$, and the mean sizes, of E, E+S0 and E+S0+Sa samples are within 10\\% of those in the Hyde \\& Bernardi (2009), $C_r\\ge 2.86$ and $C_r\\ge 2.6$ samples, respectively. Summed over all galaxy types, we find $\\rho_* \\sim 3\\times 10^8 M_\\odot$Mpc$^{-3}$ at $z\\sim 0$. This is in good agreement with expectations based on integrating the star formation history. However, compared to most previous work, we find an excess of objects at large masses, up to a factor of $\\sim 10$ at $M_*\\sim 5 \\times 10^{11}M_\\odot$. The stellar mass density further increases at large masses if we assume different IMFs for elliptical and spiral galaxies, as suggested by some recent chemical evolution models, and results in a better agreement with the dynamical mass function. We also show that the trend for ellipticity to decrease with luminosity is primarily because the E/S0 ratio increases at large $L$. However, the most massive galaxies, $M_*\\ge 5\\times 10^{11}M_\\odot$, are less concentrated and not as round as expected if one extrapolates from lower $L$, and they are not well-fit by pure deVaucouleur laws. This suggests formation histories with recent radial mergers. Finally, we show that the age-size relation is flat for ellipticals of fixed dynamical mass, but, at fixed $M_{\\rm dyn}$, S0s and Sas with large sizes tend to be younger. Hence, samples selected on the basis of color or $C_r$ will yield different scalings. Explaining this difference between E and S0 formation is a new challenge for models of early-type galaxy formation. ", "introduction": "Each galaxy has its own peculiarities. Nevertheless, even to the untrained eye, sufficiently well-resolved galaxies can be separated into three morphological types: disky spirals, bulgy ellipticals, and others which are neither. The morphological classification of galaxies is a field that is nearly one hundred years old, and sample sizes of a few thousand morphologically classified galaxies are now available (e.g. Fukugita et al. 2007; Lintott et al. 2008). However, such eyeball classifications are prohibitively expensive in the era of large scale sky surveys, which image upwards of a few million galaxies. Moreover, the morphological classification of even relatively low redshift objects from ground-based data is difficult. Thus, a number of groups have devised automated algorithms for discerning morphologies from such data (e.g. Ball et al. 2004 and references therein). In parallel, it has been recognized that relatively simple criteria, using either crude measures of the light profile (e.g. Strateva et al. 2001), the colors (e.g. Baldry et al. 2004), or some combination of photometric and spectroscopic information (Bernardi et al. 2003; Bernardi \\& Hyde 2009) allow one to separate early-type galaxies from the rest. Because they are so simple, these tend to be more widely used. The main goal of this paper is to show how samples based on such crude cuts compare with those which are based on the eyeball morphological classifications of Fukugita et al. (2007). We do so by comparing the luminosity, stellar mass, size and velocity dispersion distributions for cuts based on photometric parameters with those based on morphology. These were chosen because the luminosity function is standard, although it is becoming increasingly common to compare models with $\\phi(M_*)$ rather than $\\phi(L)$ (e.g. Cole et al. 2001; Bell et al. 2003; Panter et al. 2007; Li \\& White 2009); the size distribution $\\phi(R)$ has also begun to receive considerable attention recently (Shankar et al. 2009b); and the distribution of velocity dispersions $\\phi(\\sigma)$ (Sheth et al. 2003) is useful, amongst other things, to reconstruct the mass distribution of super-massive black holes (e.g. Shankar et al. 2004; Tundo et al. 2007; Bernardi et al. 2007; Shankar, Weinberg \\& Miralda-Escud{\\'e} 2009; Shankar et al. 2009a) and in studies of gravitational lensing (Mitchell et al. 2005). Section~\\ref{data} describes the dataset, the photometric and spectroscopic parameters derived from it, and the subsample defined by Fukugita et al. (2007). This section shows how we use quantities output from the SDSS database to define seeing-corrected half-light radii which closely approximate the result of bulge + disk decompositions. We describe our stellar mass estimator in this section as well; a detailed comparison of it with stellar mass estimates computed by three different groups is presented in Appendix~\\ref{Mscomp}. The result of classifying objects into two classes, on the basis of color, concentration index, or morphology are compared in Section~\\ref{sec:mcba}. Luminosity, stellar mass, size and velocity dispersion distributions, for the Fukugita et al. morphological types are presented in Section~\\ref{DFs}, where they are compared with those based on the other simpler selection cuts. This section includes a discussion of the functional form, a generalization of the Schechter function, which we use to fit our measurements. We find more objects with large stellar masses than in previous work (e.g. Cole et al. 2001; Bell et al. 2003; Panter et al. 2007; Li \\& White 2009); this is the subject of Section~\\ref{phiMass}, where implications for the match with the integrated star formation rate, and the question of how the most massive galaxies have evolved since $z\\sim 2$ are discussed. While we believe these distributions to be interesting in their own right, we also study a specific example in which correlations between quantities, rather than the distributions themselves, depend on morphology. This is the correlation between the half-light radius of a galaxy and the luminosity weighted age of its stellar population. Section~\\ref{ageRe} shows that the morphological dependence of this relation means it is sensitive to how the `early-type' sample was selected, potentially resolving a discrepancy in the recent literature (Shankar \\& Bernardi 2009; van der Wel et al. 2009; Shankar et al. 2009c). A final section summarizes our results, many of which are provided in tabular form in Appendix~\\ref{tables}. Except when stated otherwise, we assume a spatially flat background cosmology dominated by a cosmological constant, with parameters $(\\Omega_m,\\Omega_\\Lambda)=(0.3,0.7)$, and a Hubble constant at the present time of $H_0=70$~km~s$^{-1}$Mpc$^{-1}$. When we assume a different value for $H_0$, we write it as $H_0=100h$~km~s$^{-1}$Mpc$^{-1}$. ", "conclusions": "\\label{discuss} We compared samples selected using simple selection algorithms based on available photometric and spectroscopic information with those based on morphological information. Requiring concentration indices $C_r\\ge 2.6$ selects a mix in which E+S0+Sa's account for about two-thirds of the objects; if $C_r\\ge 2.86$ instead, then two-thirds of the sample comes from E+S0s; whereas Es alone account for more than two-thirds of a sample selected following Hyde \\& Bernardi (2009) (Figures~\\ref{fredC} and~\\ref{fblueC}, and Table~\\ref{purity}). E's alone account for about 40\\%, 50\\% and 75\\% of the total stellar mass in samples selected in these three ways. The reddest objects at intermediate luminosities or stellar masses are edge-on disks (Figure~\\ref{gmrmorph}). As a result, samples selected on the basis of color alone, or cuts which run parallel to the red sequence are badly contaminated by such objects. However, simply adding the additional requirement that the axis ratio $b/a\\ge 0.6$ is an easy way to remove such red edge-on disks from the `red' sequence; the resulting sample is similar to requiring $C_r\\ge 2.86$. This may provide a simple way to select relatively clean early-type samples in higher redshift datasets (e.g. DEEP2, $z$Cosmos). Our measurements provide the low redshift benchmarks against which such future higher redshift measurements can be compared. We showed how the distribution of luminosity, stellar mass, size and velocity dispersion in the local universe is partitioned up amongst different morphological types, and we compared these distributions with those based on simple selection algorithms based on available photometric and spectroscopic information (Figures~\\ref{differentCIa}--\\ref{ES0Sa}). We described our measurements by assuming that the intrinsic distributions have the form given by equation~(\\ref{phiX}). We showed how measurement errors bias the fitted parameters (equation~\\ref{psiO}), and used this to devise a simple method which removes this bias. The results, which are reported in tabular form in Appendix~\\ref{tables}, show that ellipticals contain $\\sim$20\\% of the luminosity density and 25\\% of the stellar mass density in the local universe, and have mean sizes of order 3.2~kpc. Including S0s increases these numbers to 33\\% and 40\\%; adding Sas results in further increases, to 50\\% and 60\\% respectively. These numbers are in broad agreement with those from the Millennium Galaxy Survey of about $10^4$ objects in $37.5$~deg$^2$. Driver et al. (2007) report that $15\\pm 5$\\% of the stellar mass density is in ellipticals, and adding bulges increases this to $44\\pm 9$\\%. Our stellar mass function has more massive objects than other recent determinations (e.g. Cole et al. 2001; Bell et al. 2003; Panter et al. 2007; Li \\& White 2009), similarly shifted to a Chabrier (2003) IMF (Figures~\\ref{MsFpet} and ~\\ref{MsFmod}). The mass scale on which the discrepancy arises is of order where some previous work had only a handful of objects -- our substantially larger volume is necessary to provide a more reliable estimate of these abundances. Using stellar masses estimated from {\\tt cmodel} luminosities, which are more reliable than Petrosian luminosities at the large masses where the discrepancy in $\\phi(M_*)$ is largest, gives stellar mass densities in objects more massive than $(1,2,3)\\times 10^{11}\\,M_\\odot$ that are larger by more than $\\sim$~$(20,50,100)$ percent compared to Bell et al. (2003) (Figure~\\ref{MsFcumulative}). This analysis required that we study the sytematic differences between the stellar mass estimates based on $g-r$ color (our equation~\\ref{gmrBell}, following Bell et al. 2003), colors in multiple bands (Blanton \\& Roweis 2007), and on spectral features (Gallazzi et al. 2005). (See Gallazzi \\& Bell 2009, which appeared while our work was being refereed, for a discussion of the pros and cons of these various approaches, and of the accuracy to which stellar masses can currently be derived.) The $g-r$ and Gallazzi et al. estimates are generally in good agreement (Figure~\\ref{BG}), although the spectral based estimates suffer slightly from aperture effects which are complicated by the magnitude limit of the survey (Figures~\\ref{BGz}, \\ref{CompareMsz} and~\\ref{MsFpet}). The Blanton et al. estimates are in good agreement with the other two provided one uses LRG-based templates to estimate masses at the most massive end (Figures~\\ref{BG-BR}, \\ref{CompareMsAll} and~\\ref{MsFmod}). At lower masses, some combination of the LRG and other templates is required. Ignoring the LRG templates altogether (e.g. Li \\& White 2009) results in systematic underestimates of as much as 0.1~dex or more (Figure~\\ref{BG-BR}), severely compromising estimates of the number of stars currently locked up in massive galaxies (Figure~\\ref{MsFpet}). If we compare our estimate of the stellar mass density in objects more massive than $(1,2,3)\\times 10^{11}\\,M_\\odot$ with those from the Li \\& White (2009) fit, then our values are $\\sim$~$(140,230,400)$ percent larger. Allowing more high mass objects means that major dry mergers may remain a viable formation mechanism at the high mass end (Figure~\\ref{Bezanson}). It also relieves the tension between estimates of the evolution of the most massive galaxies which are based on clustering (which predict some merging, and so some increase in stellar mass; Wake et al. 2008; Brown et al. 2008) and those based on abundances (for which comparison of high redshift measurements with the previous $z\\sim 0$ measurements indicated little evolution; Wake et al. 2006; Brown et al. 2007; Cool et al. 2008). This discrepancy may be related to the origin of intercluster light (e.g. Skibba et al. 2007; Bernardi 2009); our measurement of a larger local abundance in galaxies reduces the amount of stellar mass that must be stored in the ICL. It has been argued that a number of observations are better reproduced if one assumes a different IMF for elliptical and spiral galaxies (e.g. Calura et al. 2009). We showed that this acts to further increase the abundance of massive galaxies (Figure~\\ref{MsFIMF}), and reduces the difference between stellar and dynamical mass, especially at larger masses. At $M_*\\ge 10^{11}M_\\odot$, the increase due to the change in IMF is a factor of two with respect to models which assume a fixed IMF. If we sum up the observed counts to estimate the stellar mass density in the range $8.6 < \\log_{10} M_* < 12.2$ $M_{\\odot}$ ($M_*$ from equation~\\ref{gmrBell} using {\\tt cmodel} magnitudes), then the result is $3.05\\times 10^8M_\\odot$~Mpc$^{-3}$. Using our fit to the observed distribution (values between round brackets in Table~\\ref{tabMs}) gives a similar value ($3.06\\times 10^8M_\\odot$~Mpc$^{-3}$) and a slightly smaller value if one uses the {\\it intrinsic} fit ($2.89\\times 10^8M_\\odot$~Mpc$^{-3}$, see Table~\\ref{tabMs}). Our values are $\\sim 15$\\% and 30\\% larger than those reported by Panter et al. (2007) and Li \\& White (2009), respectively. If we allow a type dependent IMF, the total stellar mass density increases by a further 30\\%. However, although our stellar mass function has more $M_* > 10^{11}$ M$_\\odot$ objects than other recent determinations, our estimate of the total stellar mass density is similar to that measured by Bell et al. (2003). It is about 20\\% smaller than the value reported by Driver et al. (2007) (once shifted to the same IMF, for which we have chosen Chabrier 2003; see Table~\\ref{tabIMF}). This is because differences at the mid/faint end contribute more to the total stellar density than the difference we measured at the massive end. It has been suggested that direct integration of the cosmological star formation rate overpredicts the total local estimate of the stellar mass density (see, e.g., Wilkins et al. 2008, and references therein). However, we showed that recent determinations of the recycling factor (equation~\\ref{eq|rhoz}) and the high-$z$ star formation rate (Figure~\\ref{fig|SFRz}) result in better agreement (Figure~\\ref{fig|rhoz}). This is because the former yields smaller remaining masses, and the latter produces fewer stars formed in the first place. Our measurements also show that the most luminous or most massive galaxies, which one might identify with BCGs, are less concentrated and have smaller $b/a$ ratios, than slightly less luminous or massive objects (Figures~\\ref{cmorph}, \\ref{bamorph} and~\\ref{fredC}). Their light profile is also not well represented by a pure deVaucoleur law. This is consistent with results in Bernardi et al. (2008) and Bernardi (2009) who suggest that these are signatures of formation histories with recent radial mergers. In this context, note that we showed how to define seeing-corrected sizes, using quantities output by the SDSS pipeline, that closely approximate deVaucouleur bulge + Exponential disk decompositions (equations~\\ref{Rcmodel} and~\\ref{skysubr}). Our {\\tt cmodel} sizes represent a substantial improvement over Petrosian sizes (which are not seeing corrected) and pure {\\tt deV} or {\\tt Exp} sizes (Figures~\\ref{cmodel} and~\\ref{cmodel2}). And finally, our study of the age-size correlation resolves a discrepancy in the literature: whereas Shankar et al. (2009c) report no correlation at fixed $M_{\\rm dyn}$, van der Wel et al. (2009) report that larger galaxies tend to be younger. We showed that ellipticals follow the scaling reported by Shankar et al. scaling, whereas S0s and Sas follow that of van der Wel et al. (Figure~\\ref{ageRMd}), suggesting that Shankar et al. select a sample dominated by galaxies with elliptical morphologies, whereas van der Wel et al. include more S0s and Sas. These conclusions about the differences between the samples are consistent with how the samples were actually selected. Since van der Wel et al. use their measurements to constrain a model for early-type galaxy formation, this is an instance in which having morphological information matters greatly for the physical interpretation of the data. Our results indicate that models of early-type galaxy formation should distinguish between ellipticals and S0s because, in the projection of the age-size-mass correlation shown in Figure~\\ref{ageRMd}, the S0s and Sas are very different from the other morphological types. Whether the smaller sizes for older S0s are due to the gradual stripping away of a younger disk is an open question. van der Wel et al. (2009) use their observation that the age-size relation at fixed $\\sigma$ is flat to motivate a model in which early-type galaxy formation requires a critical velocity dispersion (which they allow to be redshift dependent). The same logic applied here suggests that while this may be reasonable for S0s or Sas, it is not well-motivated for ellipticals (Figure~\\ref{ageRsig}). However, it might be interesting to explore a model in which elliptical formation requires a critical (possibly redshift dependent) dynamical mass rather than velocity dispersion (which may be redshift dependent). This is interesting because, in hierarchical models, the phenomenon known as down-sizing (e.g., Cowie et al. 1996; Heavens et al. 2004; Sheth et al. 2006; Jimenez et al. 2007) is then easily understood (Sheth 2003)." }, "0910/0910.4480_arXiv.txt": { "abstract": "A search for muon neutrinos from Kaluza-Klein dark matter annihilations in the Sun has been performed with the 22-string configuration of the IceCube neutrino detector using data collected in 104.3 days of live-time in 2007. No excess over the expected atmospheric background has been observed. Upper limits have been obtained on the annihilation rate of captured lightest Kaluza-Klein particle (LKP) WIMPs in the Sun and converted to limits on the LKP-proton cross-sections for LKP masses in the range 250 -- 3000 GeV. These results are the most stringent limits to date on LKP annihilation in the~Sun. ", "introduction": " ", "conclusions": "" }, "0910/0910.4449_arXiv.txt": { "abstract": "{} {We investigate the temporal evolution of magnetic flux emerging within a granule in the quiet-Sun internetwork at disk center.} {We combined IR spectropolarimetry of high angular resolution performed in two \\ion{Fe}{i} lines at 1565\\,nm with speckle-reconstructed G-band imaging. We determined the magnetic field parameters by a LTE inversion of the full Stokes vector using the SIR code, and followed their evolution in time. To interpret the observations, we created a geometrical model of a rising loop in 3D. The relevant parameters of the loop were matched to the observations where possible. We then synthesized spectra from the 3D model for a comparison to the observations.} {We found signatures of magnetic flux emergence within a growing granule. In the early phases, a horizontal magnetic field with a distinct linear polarization signal dominated the emerging flux. Later on, two patches of opposite circular polarization signal appeared symmetrically on either side of the linear polarization patch, indicating a small loop-like structure. The mean magnetic flux density of this loop was roughly 450\\,G, with a total magnetic flux of around 3$\\times 10^{17}$\\,Mx. During the $\\sim$12\\,min episode of loop occurrence, the spatial extent of the loop increased from about 1 to 2\\,arcsec. The middle part of the appearing feature was blueshifted during its occurrence, supporting the scenario of an emerging loop. There is also clear evidence for the interaction of one loop footpoint with a preexisting magnetic structure of opposite polarity. The temporal evolution of the observed spectra is reproduced to first order by the spectra derived from the geometrical model. During the phase of clearest visibility of the loop in the observations, the observed and synthetic spectra match quantitatively.} {The observed event can be explained as a case of flux emergence in the shape of a small-scale loop. The fast disappearance of the loop at the end could possibly be due to magnetic reconnection.} ", "introduction": "\\label{introduction} Our knowledge of small-scale magnetic fields located in the photosphere and low chromosphere has evolved extremely fast thanks to the high spatial resolution observations done with the Hinode satellite (e.g., Lites et al. \\cite{litesetal08}), and the improvement of the spatial resolution of large ground-based solar telescopes achieved by adaptive optics systems (e.g., von der L{\\\"u}he et al. \\cite{vdlueheetal03}). The results of such observations tracing the weakest magnetic fields can now be compared directly with theoretical models, for example on the question whether a local dynamo operates in the solar photosphere (Cattaneo \\cite{cattaneo99}). The observations may also help to refine modern numerical simulations of surface magnetoconvection (V\\\"{o}gler et al. \\cite{vogleretal05}, Schaffenberger et al. \\cite{schaffenbergeretal06}, Stein \\& Nordlund \\cite{stein_nordlund06}, Abbett \\cite{abbett07}) by providing more stringent boundary conditions on magnetic field properties in the solar photosphere. Direct measurements of the 3D topology and the temporal evolution of magnetic loops associated with small granular and intergranular structures are, however, still rare. The measurements require accurate spectropolarimetric 2D data of all Stokes parameters with high spatial resolution, high signal-to-noise ratio, and high cadence. Mart\\'{\\i}nez Gonz\\'{a}lez et al. (\\cite{martinezgonzalezetal07}) reported observations of the magnetic field vector in the internetwork. They showed evidence that small-scale (only 2-6\\,arcsec long) low-lying loops connect at least 10-12\\% of the internetwork magnetic flux measured at 1565\\,nm. Centeno et al. (\\cite{centenoetal07}, CE07) and Orozco Su\\'{a}rez et al. (\\cite{orozcosuarezetal08}, OR08) analyzed the temporal evolution of such small-scale magnetic features. They used similar data sets of the \\ion{Fe}{i}\\,630.15\\,nm and 630.25\\,nm spectral lines from the Hinode/SOT spectropolarimeter (Lites et al. \\cite{litesetal01}, Kosugi et al. \\cite{kosugietal07}). OR08 % constructed maps of the temporal evolution of linear and circular polarization signals, line-of-sight velocity, and the intensities of the \\ion{Fe}{i}\\,630.25\\,nm and \\ion{Ca}{ii}\\,H spectral line. They found magnetic signals appearing at the central parts of granules, but without a significant linear polarization signal. Estimated lifetimes of the events were 20 and 14\\,min. They did not find any significant coupling of the events to the chromosphere. CE07 % applied an inversion assuming local thermodynamic equilibrium (LTE) to the Hinode spectra to retrieve information on the temporal evolution of the magnetic flux and its topology. Similar to OR08, % they also reported the appearance of magnetic signals in the central parts of granules, and documented a drift of vertical dipoles toward the surrounding intergranular lanes. CE07 found a slightly shorter lifetime of only 8 min. Moreover, they found that the appearance of horizontal magnetic fields precedes that of the vertical fields. They interpreted their results as the rise of small-scale magnetic loops through the photosphere, although they could not exclude a descending loop. In this paper, we present a similar analysis of the temporal evolution of the appearance and topology changes during a particular flux emergence event in the internetwork at the disk center, using ground-based observations in the infrared spectral region rendering high polarimetric precision and magnetic sensitivity. We compare these observations with synthetic spectra from a geometrical model of a rising loop to investigate whether the observations can be interpreted in such terms. ", "conclusions": "\\label{discus_and_concl} High-resolution surveys of large areas of quiet-Sun internetwork (Lites et al. \\cite{litesetal08}) and plage regions (Ishikawa et al. \\cite{ishikawaetal08}) revealed widespread occurrences of relatively strong horizontal fields over granules, accompanied by vertical bipolar fields cospatial with up- and downflows (Mart\\'{\\i}nez Gonz\\'{a}lez et al. \\cite{martinezgonzalezetal07}). On the other hand, individual granules are often collocated with strong unipolar fields decoupled from granular outflows (OR08, % Bello Gonz\\'{a}lez et al. \\cite{bellogonzalezetal08}). \\begin{figure} \\centerline{\\resizebox{8.8cm}{!}{\\includegraphics{small-12807f12.eps}}} \\caption{Comparison of observed ({\\em bottom}) and simulated spectra ({\\em top}). {\\em Left to right}: Stokes $IQUV$. {\\em Bottom to top}: Time steps 1 to 6. The display ranges for $QUV$ are given at {\\em bottom right} in each panel. The {\\em vertical dashed line} in Stokes $I$ denotes the rest wavelength of the 1564.8\\,nm line. The short {\\em horizontal bars} at the left in $IQUV$ of step 4 denote the location of the profiles shown in Fig.\\,\\ref{fig_profiles}. } \\label{spec_comp} \\end{figure} CE07 % reported for the first time the emergence of a magnetic loop inside of a granule whose footpoints drifted towards intergranular lanes. Here we present new observational evidence for a similar event, whose onset manifests itself through enhanced net linear polarization within an evolved granule. Soon after, a pair of opposite circularly polarized patches appear symmetrically on each side of the linear polarization forming a small magnetic dipole bridged by a horizontal magnetic field. As long as both patches of circular polarization are visible, they are about of the same size and show a balanced magnetic flux of opposite sign ($+2.9$ and $-3.1\\times 10^{17}$ Mx in step 4). This supports the conclusion made by Lamb et al. (\\cite{lambetal08}) that the model of an asymmetric dipole emergence is very unlikely. The fast disappearance of the linear polarization signal after 10:43:53\\,UT (step 5, see Figs.\\,\\ref{fig_tip_overview}, \\ref{fig_gband_magparam}, or \\ref{2d_maps}) is in striking contrast with its gradual growth seen from 10:38:47\\,UT to 10:43:53\\,UT. Can we understand it just by an upward extending loop reaching up to a higher atmospheric layer undetectable by the \\ion{Fe}{i} lines as suggested in CE07? % If we consider the loop top as a magnetic perturbation propagating upwards through otherwise non-magnetic layers, then a smooth decay of contribution and response functions with height would imply slower ceasing of the net linear polarization, perhaps detectable on the next frame. Since the contribution and response functions of Stokes {\\em Q} and {\\em U} to magnetic perturbations can be very complex (del Toro Iniesta \\cite{deltoroiniesta_book03}), their detailed computations would help to understand the disappearance of linear polarization after 10:43:53\\,UT. But after 10:43:53\\,UT the circular polarization signal of the footpoint that meets an opposite-polarity magnetic field also weakens rapidly. This could indicate magnetic reconnection between the emergent flux and the preexisting magnetic fields. If reconnection thus is such a common feature, it could explain why OR08 found no trace of the flux emergence in the chromospheric layers. The disappearance of the polarization signal could, however, also be related to other causes. In the rising loop simulation, the linear polarization amplitude at the last step is again comparable to the first one. Taking into account the noise present in the observations, the polarization signal could also be simply reduced below the detection limit. Another possibility could be a slight spatial drift of the feature with time that would remove it partly out of the observed FoV, because it already was located only in the first three steps of each repeated scan (see Fig.\\,\\ref{fig_tip_overview}). Moreover, the possible expansion of the magnetic loop with height (and time) could also affect the measurement of the linear signal above the detection threshold. From the comparison with the simulated loop, we conclude, however, that at least the ascent until maximal linear polarization signal is reached is fully traced by the observations. Our flux density estimate of $\\sim$450\\,G found in the emerging loop is comparable to the values reported in Mart\\'{\\i}nez Gonz\\'{a}lez et al. (\\cite{martinezgonzalezetal07}) and Ishikawa et al. (\\cite{ishikawaetal08}) for small-scale magnetic loops in the solar internetwork and plage region, respectively; it significantly exceeds, however, the 10-100\\,G given by Mart\\'{\\i}nez Gonz\\'{a}lez \\& Bellot Rubio (\\cite{martinezgonzalez_bellotrubio09}). This could be due to the assumption of a magnetic filling factor of unity by the latter authors. The loop footpoints separate with time by around 2\\,km\\,s$^{-1}$, drifting in opposite directions towards intergranular lanes. Moreover, our estimate of the magnetic flux density shows that the field is stronger than the typical equipartition value of 400\\,G (Cheung et al. \\cite{cheungetal07}, Ishikawa et al. \\cite{ishikawaetal08}). This suggests that both granular outflow and relaxation of the magnetic tension of the loop may drive the footpoints. How does our loop compare to the one reported in CE07? % They occupy similar locations and enclose too little magnetic flux to leave any trace on the underlying granules or ambient G-band bright points. On the other hand, our loop displays a higher magnetic flux density and a longer lifetime. \\begin{figure} \\resizebox{8.8cm}{!}{\\includegraphics{small-12807f13.eps}} \\caption{Comparison of observed ({\\em black}) and synthetic spectra ({\\em gray}) in step 4. {\\em Top to bottom}: Stokes $IQUV$. {\\em Left to right}: Lower footpoint, center of loop, upper footpoint. } \\label{fig_profiles} \\end{figure} Another difference to the loop reported in CE07 % is that in our case one loop footpoint meets the ambient field of opposite polarity, which is followed by the fast disappearance of the loop, in contrast to its more gradual emergence. This is very symptomatic of field cancellation by magnetic reconnection, as reported in Cheung et al (\\cite{cheungetal08}) and Guglielmino et al. (\\cite{guglielminoetal08}). The magnetic reconnection likely drags loop remnants very rapidly out of the line-forming domain less than 12 min after the first trace of the loop top was detected. The magnetic loop emergence presented here is probably just a particular event in a multitude of similar ones occurring continuously in the solar internetwork. Simulations of small-scale magnetoconvection by, e.g., Stein \\& Nordlund (\\cite{stein_nordlund06}) show examples of granular-sized magnetic loops with similar behavior. They probably build up a sub-canopy of horizontal fields over granules (Wedemeyer-B\\\"{o}hm et al. \\cite{wedemeyerbohmetal09}). Surface dynamo simulations by Sch\\\"{u}ssler \\& V\\\"{o}gler (\\cite{schussler_vogler08}) and Steiner et al. (\\cite{steineretal08}) clearly show that this new component of the internetwork is a direct consequence of magnetic flux expulsion from the granular interior, not only to the intergranular lanes, but also to the upper photosphere. Cheung et al. (\\cite{cheungetal07}, \\cite{cheungetal08}) found in numerical simulations that emerging flux regions also contribute to the flux accumulation and its maintenance above granular upflows in the magnetic inversion layer. Our results show one particular example of this local process albeit terminated, probably due to magnetic reconnection. Better temporal and spatial resolution is needed for more detailed information on the process than our observations provide." }, "0910/0910.3743_arXiv.txt": { "abstract": "{Spectroscopy of redshifted radio absorption of atomic and molecular species provide excellent probes of the cold component of the gas in the early Universe which can be used to address many important issues, such as measuring baryonic content, probing large-scale structure and galaxy evolution, as well as obtaining independent measurements of various combinations of fundamental constants at large look-back times. However, such systems are currently very rare with only 80 detected in \\HI\\ 21-cm and five in OH and millimetre-band species. Here we summarise the main selection criteria responsible for this and how the next generation of radio telescopes are expected to circumvent these through their wide instantaneous bandwidths and fields-of-view. Specifically: \\begin{enumerate} \\item {\\em \\HI\\ in absorbers occulting distant quasars:} These are usually found in known optical absorbers and wideband radio surveys could reveal a much fainter population. However, the 21-cm absorption strength may be correlated with the width of the singly ionised metal species, suggesting that these may be weak, and due to the effects of a flat expanding Universe on the covering factor, we expect the highest detection rates at $z < 1$. \\item {\\em \\HI\\ absorption associated with the host galaxy/quasar:} Due to high degrees of ionisation/excitation rendering 21-cm undetectable near active nuclei with ultra-violet luminosities of $L_{\\rm UV}\\gapp10^{23}$ \\WpHz, future searches should be magnitude limited, e.g. at $z>1$, blue magnitudes should be $B \\gapp19$ with $z>2 \\Rightarrow B \\gapp21$ and $z>3 \\Rightarrow B \\gapp22$. \\item {\\em OH (and millimetre-band) absorption:} For all of the known redshifted molecular absorption systems a correlation is found between the molecular fraction and the optical--near-infrared colour ($V-K$), with the five known OH absorbers all having $V-K\\gapp5$. Therefore spectral scans towards extremely red radio sources are expected to uncover any dusty intervening, molecular rich absorbers reponsible for the obscuration of the optical light. \\end{enumerate} } \\FullConference{Panoramic Radio Astronomy: Wide-field 1-2 GHz research on galaxy evolution\\\\ June 2-5 2009\\\\ Groningen, the Netherlands} \\begin{document} ", "introduction": " ", "conclusions": "" }, "0910/0910.0786.txt": { "abstract": "A filament disappearance event was observed on 22 May 2008 during our recent campaign JOP 178. The filament, situated in the southern hemisphere, showed sinistral chirality consistent with the hemispheric rule. The event was well observed by several observatories in particular by THEMIS. One day before the disappearance, H$\\alpha$ observations showed up and down flows in adjacent locations along the filament, which suggest plasma motions along twisted flux rope. THEMIS and GONG observations show shearing photospheric motions leading to magnetic flux canceling around barbs. STEREO A, B spacecraft with separation angle 52.4 degrees, showed quite different views of this untwisting flux rope in He II 304 \\AA\\ images. Here, we reconstruct the 3D geometry of the filament during its eruption phase using STEREO EUV He II 304 \\AA\\ images and find that the filament was highly inclined to the solar normal. The He II 304 \\AA\\ movies show individual threads, which oscillate and rise to an altitude of about 120 Mm with apparent velocities of about 100 km s$^{-1}$, during the rapid evolution phase. Finally, as the flux rope expands into the corona, the filament disappears by becoming optically thin to undetectable levels. No CME was detected by STEREO, only a faint CME was recorded by LASCO at the beginning of the disappearance phase at 02:00 UT, which could be due to partial filament eruption. Further, STEREO Fe XII 195 \\AA\\ images showed bright loops beneath the filament prior to the disappearance phase, suggesting magnetic reconnection below the flux rope. ", "introduction": "\\label{S-Introduction} Two types of structures are commonly used to model filaments i.e., arcades structure or twisted flux tubes. Evidence of twisted flux tubes clearly appear during the eruptive phase of prominences \\cite{Gary2004}, \\cite{Torok2005}. Most quantitative models assume that the flux rope is in static equilibrium. 3-D magnetic models of filaments by extrapolating photospheric magnetograms into the corona have been developed \\cite{1998A&A...329.1125A}, \\cite{Aulanier2000}, \\cite{Dudik2008}, \\cite{aad2004}. Such models reproduce helical ropes overlying the polarity inversion line (PIL) and the filament plasma is assumed to be located in dips of the helical field lines. In areas where magnetic parasitic polarity elements are located close to the PIL, the dips extend away from the main body of the filament creating barbs. It is consistent with the findings of the observers \\cite{Martin94},\\cite{Martin98} and \\cite{Wang2001}. One of the possible causes of filament eruption is instability %Instabilities in the filament equilibrium sometimes lead to eruption \\cite{Forbes1991},\\cite{Isenberg1993}. Determination of true filament height is an important parameter in studies of filament eruption \\cite{Schrijver2008}. Traditionally, this has been a difficult task and only via H$\\alpha$ observations of the limb prominence one could measure the filament height. The disadvantage of this method being: (i) it measures only the projected height, (ii) non-availability of magnetograms due to its location at the limb, (iii) can not be used to follow height evolution for several days and (iv) no continuity of multi-temperature observations. Only recently, with the advent of stereoscopic observations by the twin spacecraft of the {\\it Solar Terrestrial Relations Observatory} (STEREO) mission called STEREO-A (Ahead) and STEREO-B (Behind), the true height of filament can be judged properly and the heating of the plasma can be tested \\cite{2008SSRv..136....5K};\\cite{2008SoPh..252..397G};\\cite{Liewer09}. Here in this paper, we report on the multi-wavelength observations of a filament eruption event using ground as well as space based observatories during a joint observing campaign (JOP-178 from 20 to 25 May 2008). The filament was located in a large filament channel with very weak and diffuse polarities. The weak magnetic polarities in the filament channel are recognized using THEMIS/MTR instrument which has high polarimetric sensitivity and a simultaneous H$\\alpha$ scan. The evolution of these polarities is studied using full-disk GONG line-of-sight magnetograms which are available at a cadence of one minute. Further, we reconstruct the true filament height and eruption velocity using stereoscopic images by SECCHI/EUV instrument in He II 304 \\AA\\ ($\\sim$60-80 $\\times 10^{3}$ K). The He II 304 \\AA\\ observations are very useful in tracing filaments because (i) the filament spine is much sharper and clearer \\cite{2007AAS...21012006M};\\cite{2007BASI...35..447J}, and (ii) the filament can be traced up to higher altitudes compared to H$\\alpha$ images \\cite{2007BASI...35..447J}. With the help of stereoscopic observations by STEREO mission we reconstruct the filament geometry and height during the phase. It is difficult to derive the rise velocity by height-time profile as the filament becomes very diffuse during the disappearance phase, however, we try to estimate the projected rise velocity of individual threads from movies. Also, we identify possible reconnected loops below the flux rope. ", "conclusions": "A long S-shaped filament composed of several segments and located in the Southern hemisphere was disappearing between May 20-22, 2008. This was observed during a coordinated campaign (JOP-178) involving ground-based instruments, THEMIS on the Canary Islands, MSDP at the Meudon solar tower, GONG magnetograph and H$\\alpha$ telescope in Udaipur, as well as space-based instruments SOHO/MDI and SECCHI/EUVI aboard STEREO. H$\\alpha$ instruments observed the progressive disappearance of the filament, segment after segment, between May 20 to May 22. It was a long process with some impulsive phases. The last H$\\alpha$ segment was visible on May 22 till 10:00 UT while the long spine of the S-shaped filament was observable in He II 304 \\AA\\ till 15:30 UT. Before disappearing filament segments showed high dynamics in H$\\alpha$ with blue and red shifts parallel to the filament axis. MDI and GONG magnetograms show that the filament is located along the polarity inversion line between two weak magnetic field regions. The magnetic field in the filament channel was weak and rapidly changing. Local correlation tracking techniques applied to GONG polarities showed strong shear between these two regions two days before the eruption, then some vortex pattern and no noticeable motion after the eruption. THEMIS magnetograms allowed us to identify weak minority polarities associated with the filament feet. The canceling flux/polarities in the vicinity of the feet were observed in the GONG movies a few hours before the disappearance of the associated segment. Such cancelation of flux and disappearance of the foot-point barbs where the filament is tied to the photosphere could lead to eruptions (Raadu {\\it et al.} 1987,1988). The angular separation between STEREO A and STEREO B spacecraft was 52.4 degrees during our observations, which gave quite different views of the filament. We used different methods to study the filament disappearance using He II 304 filtergrams: (i) study of filament dynamics and estimation of the projected rise velocity of thread-like structures using He II 304 movies, (ii) by overlaying a spherical grid over the solar surface drawn over a sphere of 700 Mm to identify elevated features and filament feet and put an upper limit on the filament inclination, and (iii) using the triangulation algorithm SCC\\_MEASURE (part of STEREO data analysis library in Solarsoft package) to compute the altitude and inclination of the filament over the chromosphere during its disappearance phase. These methods give consistent results. The stereoscopic reconstructions using SECCHI/EUVI observations of 19 May 2007 filament eruption event were reported recently by \\cite{Liewer09} and \\cite{2008SoPh..252..397G}. While \\inlinecite{Liewer09} used scc\\_measure for reconstruction, \\inlinecite{2008SoPh..252..397G} used optical flow method to find displacements between features in stereoscopic pairs of SECCHI/EUVI 304 \\AA\\ images. The results of these two techniques are in good agreement \\cite{Liewer09}. These results for the 19 May 2007 event showed that the filament eruption was asymmetric and whip-like. The filament disappearance in our case is different from their study in two ways (i) filament was rooted in a weak diffuse bipolar magnetic region not in an active region, and (ii) there was no CME recorded for our filament disappearance event (however, one cannot rule out a CME as it could be below detection threshold). The length of the filament is too large spanning more than 10$^\\circ$ in latitude, as seen in He II 304 \\AA\\ images. We also use scc\\_measure to derive three-dimensional geometry of the filament which is used to derive its height and inclination. The filament was highly inclined to solar normal by about 47$^\\circ$. Such large inclinations are interesting as theoretical models of filaments consider filament sheet as thin vertical slab. The maximum filament height determined during the disappearance phase around 10:56 UT is estimated to be about 123 Mm in the middle portion of the filament. Determining height evolution of filament during the disappearance phase was difficult as it became quite diffuse, making identification of common features very difficult. The presence of fine threads running parallel to filament spine could be seen in movies. These threads are so tiny that it was difficult to track them over the chromosphere in the later phases, i.e., after around 11:00 UT. Only the EUVI movies allowed us to distinguish them over the chromosphere. The initial dense filament became an untwisting flux rope with multiple threads with a fan-shaped structure which rose and disappeared one by one into the corona. The plasma becomes optically so thin as the flux rope rapidly expands in the corona that it remains no longer visible. Bright structure observed in STEREO A and B images at 195 \\AA\\ is interpreted as reconnection loop system, located below the filament (flux-rope). This bright structure is not visible earlier at 00:00 UT and appears during the onset of rapid disappearance phase at about 06:00 UT. The plasma of the filament is less and less dense as the flux rope rises and expands. As the plasma is dispersed it is no more visible in filter 304 \\AA\\ nor in 195 \\AA\\.~ No CME was reported during this last phase (LASCO was not observing) by COR1 and COR2, the two coronagraphs aboard STEREO. The CME might be too faint to be detected or the magnetic field might be too weak to prevent the plasma's expansion to undetectable density levels. Ground based data are very useful to understand the context of the event. Up to now we have the knowledge of eruption phenomena principally by using instruments with low temporal and/or spatial resolution. These observations are still largely used to study the association of filament eruption and CMEs. On the other hand high-cadence instrumentation with high-resolution have made a breakthrough in the nature of the dynamics of filament fine-structures. Non-eruptive filaments show counter-streaming flows along the fine structures and up-and-down flows in the filament-end (\\cite{Zirker98}, \\cite{Lin2005}, \\cite{Berger2008}). These velocities are lower than 10 km s$^{-1}$. With STEREO and ground-based observations we can find some relationship between the dynamics at large-scale and at small-scale. The H$\\alpha$ doppler velocities associated with the filament, that we measure in the present paper, are relatively small but they confirm previous observations of activated filaments with low spatial-resolution instruments (\\cite{Schmieder85a},\\cite{Schmieder85b}). The pair of aligned and elongated regions of oppositely directed velocities are interpreted in terms of a twisted magnetic flux rope. The STEREO observations allow perspective effect to be removed in doppler velocities by taking into account the filament inclination. The corrected velocities are about 1.5 times more. The Doppler-shifts are also lower due to the computation by using the bisector method and not taking account the background chromosphere. In the paper by \\cite{Schmieder85a} and \\cite{Schmieder85b}, they could derive larger velocities by a factor 2 between the foot-points and of the order of 10 km s$^{-1}$ in the feet using a cloud model method. That is probably what we could expect with our present observations using the cloud model method. The standard method (bisector) indicates the general trend of the velocities with large uncertainty but indicates clearly that the filament is activated. The day before we did not observe such organized velocity pattern. In the future, development of ground based instruments like dual-beam doppler imaging system, being developed at Udaipur Solar Observatory \\cite{Joshi2009} for detecting filament activation using high-cadence H$\\alpha$ dopplergrams, and combined observations with STEREO will lead to further developments in our understanding of the filament eruptions. \\begin{acks} The authors thank CNRS for allocating observing time on THEMIS. THEMIS is a French-Italian telescope installed at Observatorio El Teide, Tenerife, Spain. The observations from BBSO, GONG and SOHO are acknowledged. We thank V. Bommier for providing her UNNOFIT code for inversion of Stokes profiles. We also thank Arturo Lopez Ariste, G. Ruyman and Cyril Delaigue for help in observations and data reductions. Further, SG acknowledges CEFIPRA funding for his visit to Observatoire de Paris, Meudon, France under its project No. 3704-1. Also, SG acknowledges travel support to THEMIS, Tenerife by Observatoire de Paris, Meudon. We thank Fr$\\acute{\\rm e}$d$\\acute{\\rm e}$rique Auch$\\grave{\\rm e}$re for the STEREO movies. We also thank G. Molodij, J. Moity for the observations at the solar tower and P. Mein for helping in the MSDP data reduction. This work was supported by the European network SOLAIRE (MTRN\\_CT\\_2006\\_035484). \\end{acks}" }, "0910/0910.5023_arXiv.txt": { "abstract": "Measurements of small-scale turbulent fluctuations in the solar wind find a non-zero right-handed magnetic helicity. This has been interpreted as evidence for ion cyclotron damping. However, theoretical and empirical evidence suggests that the majority of the energy in solar wind turbulence resides in low frequency anisotropic kinetic \\Alfven wave fluctuations that are not subject to ion cyclotron damping. We demonstrate that a dissipation range comprised of kinetic \\Alfven waves also produces a net right-handed fluctuating magnetic helicity signature consistent with observations. Thus, the observed magnetic helicity signature does not necessarily imply that ion cyclotron damping is energetically important in the solar wind. ", "introduction": "The identification of the physical mechanisms responsible for the dissipation of turbulence in the solar wind, and for the resulting heating of the solar wind plasma, remains an important and unsolved problem of heliospheric physics. An important clue to this problem is the observed non-zero fluctuating magnetic helicity signature at scales corresponding to the dissipation range of solar wind turbulence. \\citet{Matthaeus:1982a} first proposed the ``fluctuating'' magnetic helicity as a diagnostic of solar wind turbulence, defining the ``reduced fluctuating'' magnetic helicity spectrum derivable from observational data (see \\S \\ref{sec:red} below). A subsequent study, corresponding to scales within the inertial range, found values that fluctuated randomly in sign, and suggested an interpretation that ``a substantial degree of helicity or circular polarization exists throughout the wavenumber spectrum, but the sense of polarization or handedness alternates randomly'' \\citep{Matthaeus:1982b}. Based on a study of the fluctuating magnetic helicity of solutions to the linear Vlasov-Maxwell dispersion relation, \\citet{Gary:1986} suggested instead that, at inertial range scales, all eigenmodes have a very small {\\it intrinsic} normalized fluctuating magnetic helicity, eliminating the need to invoke an ensemble of waves with both left- and right-handed helicity to explain the observations. Subsequent higher time resolution measurements, corresponding to scales in the dissipation range, exhibited a non-zero net reduced fluctuating magnetic helicity signature, with the sign apparently correlated with the direction of the magnetic sector \\citep{Goldstein:1994}. Assuming dominantly anti-sunward propagating waves, the study concluded that these fluctuations had right-handed helicity. The proposed interpretation was that left-hand polarized Alfv\\'en/ion cyclotron waves were preferentially damped by cyclotron resonance with the ions, leaving undamped right-hand polarized fast/whistler waves as the dominant wave mode in the dissipation range, producing the measured net reduced fluctuating magnetic helicity. We refer to this as the \\emph{cyclotron damping interpretation}. A subsequent analysis of more solar wind intervals confirmed these findings for the dissipation range \\citep{Leamon:1998a}. \\citet{Leamon:1998b} argued that a comparison of the normalized cross-helicity in the inertial range (as a proxy for the dominant wave propagation direction in the dissipation range) to the measured normalized reduced fluctuating magnetic helicity provides evidence for the importance of ion cyclotron damping, which would selectively remove the left-hand polarized waves from the turbulence; using a simple rate balance calculation, they concluded that the ratio of damping by cyclotron resonant to non-cyclotron resonant dissipation mechanisms was of order unity. A recent study performing the same analysis on a much larger data set concurred with this conclusion \\citep{Hamilton:2008}. In this Letter, we demonstrate that a dissipation range comprised of kinetic \\Alfven waves produces a reduced fluctuating magnetic helicity signature consistent with observations. A dissipation range of this nature results from an anisotropic cascade to high perpendicular wavenumber with $k_\\perp \\gg k_\\parallel$; such a cascade is consistent with existing theories for low-frequency plasma turbulence \\citep{Goldreich:1995,Boldyrev:2006,Howes:2008a,Schekochihin:2009}, numerical simulations \\citep{Cho:2000,Howes:2008b}, and observations in the solar wind \\citep{Horbury:2008,Podesta:2009a}. Our results imply that no conclusions can be drawn about the importance of ion cyclotron damping in the solar wind based on the observed magnetic helicity signature alone. ", "conclusions": "\\label{sec:discuss} Predicting the values of $H_m^{'r}(\\omega')$ for solar wind turbulence based on equation (\\ref{eq:hmr}) requires understanding three issues: the scaling of the magnetic fluctuation spectrum with wavenumber, the imbalance of \\Alfven wave energy fluxes in opposite directions along the mean magnetic field, and the variation of the angle $\\theta$ between the solar wind velocity $\\V{v}$ and the mean magnetic field. The 1-D magnetic energy spectrum in the solar wind typically scales as $k_1^{-5/3}$ in the inertial range and $k_1^p$ in the dissipation range, where $-2\\le p \\le -4$ \\citep{Smith:2006} and the effective wavenumber is $k_1=\\omega'/v$. It is clear from \\eqref{eq:hmr} that, when the plasma frame frequency $\\omega$ is negligible, the Doppler-shifted observed frequency always results in an effective wavenumber $k_1 \\le k$, with equality occurring only when the velocity $\\V{v}$ is aligned with the wave vector $\\V{k}$. We assume that, for homogeneous turbulence at the dissipation range scales, turbulent energy at fixed $k_\\perp$ and $k_\\parallel$ is uniformly spread over wave vectors with all possible angles $\\alpha$ about the mean magnetic field. Because the fluctuation amplitude deceases for larger effective wavenumbers, the contribution to $H_m^{'r}(\\omega')$ is maximum at angle $\\alpha=0$; for angles $\\alpha$ yielding a Doppler shift to lower effective wavenumbers $k_1<(k_\\perp^2+k_\\parallel^2)^{1/2}$, the higher amplitude fluctuations at those lower wavenumbers will contribute more strongly to $H_m^{'r}(\\omega')$. An accurate calculation of the magnetic helicity signature based on \\eqref{eq:hmr} must take into account the scaling of the magnetic energy spectrum. \\begin{figure} \\resizebox{3.1in}{!}{\\includegraphics*[0.3in,2.in][8.0in,5.1in]{f2.eps}} \\caption{\\label{fig:mhelr} Normalized reduced fluctuating magnetic helicity $\\hat{\\sigma}^r_m(k_1)$ vs.~effective wavenumber $k_1$ due to a turbulent spectrum of kinetic \\Alfven waves with $\\theta=60^\\circ$. The solid line corresponds to the model 1-D energy spectrum while the dashed line corresponds to a $k^{-1}$ spectrum.} \\end{figure} To compare to $\\sigma^r_m(k_1)$ derived from observations (for example, see Figure~1 of \\citet{Leamon:1998a}), we construct the normalized quantity \\begin{equation} \\hat{\\sigma}^r_m(k_1)= \\frac{\\sum_{\\V{k}} H'_m(\\V{k})\\frac{\\V{k}'\\cdot \\V{v} } {\\V{k}'\\cdot \\V{v}+\\omega} \\delta[\\omega'- (\\V{k}'\\cdot \\V{v}+\\omega)]} { \\sum_{\\V{k}} [|\\V{B}(\\V{k})|^2/k ]\\delta[\\omega'- (\\V{k}'\\cdot \\V{v}+\\omega)]}. \\label{eq:numsigm} \\end{equation} In evaluating \\eqref{eq:numsigm}, we assume a model 1-D energy spectrum\\footnote{On $150 \\times 150$ logarithmic gridpoints over $k_\\perp \\rho_i, k_\\parallel \\rho_i \\in [10^{-3},10^2]$, the model weights $B^2$ as a function of $k=({k_\\perp^2+k_\\parallel^2})^{1/2}$ using $B^2(k) = B_0^2 \\{[ (k \\rho_i)^{-1/3} + (k \\rho_i)^{4/3} ]/[1+(k \\rho_i)^{2}]\\}^2$.} that scales as $k^{-5/3}$ for $k\\rho_i \\ll 1$ and $ k^{-7/3}$ for $k\\rho_i \\gg 1$, consistent with theories for critically balanced turbulence \\citep{Goldreich:1995,Howes:2008b,Schekochihin:2009} and solar wind observations \\citep{Smith:2006}. In \\figref{fig:mhelr}, we plot $\\hat{\\sigma}^r_m(k_1)$ vs.~effective wavenumber $k_1=\\omega'/v$ for a turbulent spectrum filling the MHD \\Alfven and kinetic \\Alfven wave regimes ($k_\\perp>k_\\parallel$ and $k_\\parallel \\rho_i <1$) for $\\beta_i=1$, $T_i/T_e=1$, $v_{th_i}/c=10^{-4}$, $\\theta=60^\\circ$, and $v/v_A=10$. The contributions to $\\hat{\\sigma}^r_m(k_1)$ for all angles $\\alpha$ of each wave vector are collected in 120 logarithmically spaced bins in Doppler-shifted frequency. The results are rather insensitive to the scaling of the 1-D magnetic energy spectrum over the range from $k^{-1}$ to $k^{-4}$. The solid line in \\figref{fig:mhelr} corresponds to the model spectrum assumed above, while the dashed line corresponds to a $k^{-1}$ energy spectrum. \\figref{fig:mhelr} demonstrates that turbulence consisting of \\Alfven and kinetic \\Alfven waves produces a positive (right-handed) magnetic helicity signature in the dissipation range at $k_1 \\rho_i \\gtrsim 1$. The analysis presented in \\figref{fig:mhelr} considers only waves with $k_\\parallel >0$, so all of the waves in the summation in \\eqref{eq:hmr} are traveling in the same direction. If there were an equal \\Alfven wave energy flux in the opposite direction---a case of balanced energy fluxes, or zero cross helicity---the net $\\hat{\\sigma}^r_m(k_1)$ would be zero due to the odd symmetry of $H'_m(\\V{k})$ in $k_\\parallel$. It is often observed, at scales corresponding to the inertial range, that the energy flux in the anti-sunward direction dominates, leading to a large normalized cross helicity \\citep{Leamon:1998b}. If this imbalance of energy fluxes persists to the smaller scales associated with the dissipation range, a non-zero value of $\\hat{\\sigma}^r_m(k_1)$ is expected. However, theories of imbalanced MHD turbulence \\citep[][ and references therein]{Chandran:2008} predict that the turbulence is ``pinned'' to equal values of the oppositely directed energy fluxes at the dissipation scale. This implies that, at sufficiently high wavenumber $k_1$, the value of $\\hat{\\sigma}^r_m(k_1)$ should asymptote to zero. Thus, $\\hat{\\sigma}^r_m(k_1)$ in \\figref{fig:mhelr} would likely drop to zero more rapidly than shown, leaving a smaller positive net $\\hat{\\sigma}^r_m(k_1)$ around $k_1 \\rho_i \\sim 1$, consistent with observations \\citep{Goldstein:1994,Leamon:1998a,Hamilton:2008}. We defer a detailed calculation of the effects of imbalance to a future paper. The angle $\\theta$ between $\\V{B}_0$ and $\\V{v}$ is likely to vary during a measurement; this angle does not typically sample its full range $0 \\le \\theta \\le \\pi$, but has some distribution about the Parker spiral value. Calculations of $\\hat{\\sigma}^r_m(k_1)$ over $0 \\le \\theta \\le \\pi$ yield results that are qualitatively similar to \\figref{fig:mhelr}, so this averaging will not significantly change our results. Taken together, we have demonstrated that a solar wind dissipation range comprised of kinetic \\Alfven waves produces a magnetic helicity signature consistent with observations, as presented in \\figref{fig:mhelr}. The underlying assumption of the cyclotron damping interpretation of magnetic helicity measurements, an interpretation that dominates the solar wind literature \\citep{Goldstein:1994,Leamon:1998b,Leamon:1998a,Hamilton:2008}, is the slab model, $\\V{k}=k_\\parallel \\zhat$ and $k_\\perp=0$, i.e., purely parallel wave vectors. As shown in \\figref{fig:mhel}, only in the limit $k_\\parallel \\gg k_\\perp$ does the \\Alfven wave root generate a left-handed helicity $\\sigma_m \\rightarrow -1$ as $k_\\parallel \\rho_i \\rightarrow \\sqrt{\\beta_i}$; in the same limit, the fast/whistler root generates a right-handed helicity $\\sigma_m \\rightarrow +1$ in a quantitatively similar manner (see Figure~9 of \\cite{Gary:1986}). Strong ion cyclotron damping of the Alfv\\'en/ion cyclotron waves as $k_\\parallel \\rho_i \\rightarrow 1$ \\citep{Gary:2004} would leave a remaining spectrum of right-handed fast/whistler waves, as proposed by cyclotron damping interpretation. However, only if the majority of the turbulent fluctuations have $k_\\parallel \\gtrsim k_\\perp$ is the slab limit applicable, and only if significant energy resides in slab-like fluctuations are the conclusions drawn about the importance of cyclotron damping valid. There is, on the other hand, strong theoretical and empirical support for the hypothesis that the majority of the energy in solar wind turbulence has $k_\\perp \\gg k_\\parallel$ (see \\citealt{Howes:2008b} and references therein). In this case, there is a transition to kinetic \\Alfven wave fluctuations at the scale of the ion Larmor radius. This Letter demonstrates that a dissipation range comprised of kinetic \\Alfven waves produces a reduced fluctuating magnetic helicity signature consistent with observations." }, "0910/0910.2617_arXiv.txt": { "abstract": "The {\\it Chandra} COSMOS Survey (C-COSMOS) is a large, 1.8 Ms, {\\it Chandra} program, that covers the central contiguous $\\sim$~0.92 deg$^2$ of the COSMOS field. C-COSMOS is the result of a complex tiling, with every position being observed in up to six overlapping pointings (four overlapping pointings in most of the central $\\sim$~0.45 deg$^2$ area with the best exposure, and two overlapping pointings in most of the surrounding area, covering an additional $\\sim$~0.47 deg$^2$). Therefore, the full exploitation of the C-COSMOS data requires a dedicated and accurate analysis focused on three main issues: 1) maximizing the sensitivity when the PSF changes strongly among different observations of the same source (from $\\sim$~1 arcsec up to $\\sim~10$ arcsec half power radius); 2) resolving close pairs; and 3) obtaining the best source localization and count rate. We present here our treatment of four key analysis items: source detection, localization, photometry, and survey sensitivity. Our final procedure consists of a two step procedure: (1) a wavelet detection algorithm, to find source candidates, (2) a maximum likelihood Point Spread Function fitting algorithm to evaluate the source count rates and the probability that each source candidate is a fluctuation of the background. We discuss the main characteristics of this procedure, that was the result of detailed comparisons between different detection algorithms and photometry tools, calibrated with extensive and dedicated simulations. ", "introduction": "It is well known that X-ray surveys are an extremely efficient tool to select Active Galactic Nuclei (AGN). For example in the {\\it XMM-Newton} COSMOS survey, at the 0.5-2 keV limiting flux of 7$\\cdot$10$^{-16}$ erg s$^{-1}$ cm$^{-2}$, the AGN surface density is $\\sim$1000 deg$^{-2}$ (Hasinger et al. 2007, Cappelluti et al. 2007), a factor 2-4 greater than the AGN surface density in the most recent deep optical surveys, 250 deg$^{-2}$ in the COMBO-17 ( Wolf et al. 2003) and 470 deg$^{-2}$ in VVDS Survey (Gavignaud et al. 2006). There are four main causes for the higher efficiency of X-ray surveys in finding AGN: 1) X-rays directly trace the super massive black hole (SMBH) accretion, while AGN classification trough optical line spectroscopy may suffer of uncompleteness and/or misidentifications; 2) AGN are the dominant X-ray population. In fact most ($\\sim$ 80\\%) of the X-ray sources AGN in deep and shallow surveys turn out to be AGN, unlike at optical wavelengths. 3) 0.5-10 keV X-rays (the typical {\\it Chandra} and {\\it XMM-Newton} enery band) are capable to penetrate column densities up to $\\sim$10$^{24}$ cm$^{-2}$, allowing the selection of moderately obscured AGN; 4) low luminosity AGN are difficult to select in optical surveys, because their light is diluted in the host galaxy emission. So far {\\it Chandra} and {\\it XMM-Newton} have performed several deep, pencil beam, and shallower but wider surveys. Fig. \\ref{surveys} compares the flux limit and area coverage of the main {\\it Chandra} and {\\it XMM-Newton} surveys. This figure shows that {\\it XMM-Newton} COSMOS and {\\it Chandra}-COSMOS (C-COSMOS, Elvis et al. 2009, Paper I hereafter) surveys are the deepest surveys on large contiguous area. The coverage of larger areas at similar flux limits is today achieved only by serendipitous surveys using mostly not contiguous areas (see e.g., CHAMP, Kim et al. 2004a, 2004b, Green et al. 2004). \\begin{figure} \\begin{center} \\includegraphics[angle=0,height=8truecm]{xraysurveys.ps} \\caption{The 0.5-2 keV flux range vs. the area coverage for various surveys. The black solid lines represent few serendipitous surveys: Helllas2XMM (Baldi et al. 2002, symbol A), CHAMP (Kim et al. 2004a, 2004b, Green et al. 2004, symbol B), SEXSI (Harrison et al. 2003, symbol C), XMM-BSS (Della Ceca et al. 2004, symbol D), AXIS (Carrera et al. 2007, symbol E); the red dotted lines represent few deep pencil beam surveys: CDFN (Brandt et al. 2001, Alexander et al. 2003, symbol F), CDFS (Giacconi et al. 2001, Luo et al. 2008, symbol G), XMM-Newton Lockman Hole (Worsley et al. 2004, Brunner et al. 2008, symbol H); the blue dotted lines represent few wide shallow contiguous surveys: C-COSMOS (Elvis et al. 2009, symbol I), XMM-COSMOS (Hasinger et al. 2007, Cappelluti et al. 2007, 2009, symbol L), ELAIS-S1 (Puccetti et al. 2006, symbol M), E-CDF-S (Lehmer et al. 2005, symbol N), AEGIS-X (Laird et al. 2009, symbol O), SXDS (Ueda et al. 2008, symbol P). The black solid triangle represent the ROSAT all sky survey (RASS, Voges et al. 1999).} \\label{surveys} \\end{center} \\end{figure} The Cosmic evolution survey (COSMOS, Scoville et al. 2007) is aimed at studying the interplay between the Large Scale Structure (LSS) in the Universe and the formation of galaxies, dark matter, and AGN. The COSMOS field is located near the equator (10h,+02degrees), covers $\\sim$~2 square degrees as originally defined by the HST/ACS imaging (Koekemoer et al. 2007), with subsequent deep and extended multi-wavelength coverage overlapping this area. The size of COSMOS was chosen to sample LSS up to a linear size of about 50 Mpc h$^{-1}$ at z~$\\sim$~1-2, where AGN and star formation in galaxies are expected to peak. To study the role of AGN in galaxy evolution the X-ray data are fundamental. Therefore, the central square degree of the COSMOS field has been the target of a {\\it Chandra} ACIS-I, 1.8~Msec Very Large Program: the {\\it Chandra}-COSMOS survey. The C-COSMOS survey has a rather uniform effective exposure of $\\sim~160$~ksec over a large area ($\\sim$~0.45 deg$^2$), thus reaching $\\sim$~3.5 times fainter fluxes than XMM-COSMOS in both 0.5-2 keV band and 2-7 keV band. This flux limit is below the threshold where starburst galaxies become common in X-rays. The sharp {\\it Chandra} Point Spread Function (PSF) allows nearly unambiguous identification of optical counterparts (Civano et al. 2009, hereafter Paper III).{\\it Chandra} secures the identifications of X-ray sources down to faint optical magnitude (i.e., I~$\\sim$~26), with only $\\sim$~2\\% ambiguous identifications, significantly better than the $\\sim20\\%$ ambiguous identifications in XMM-Newton (Brusa et al. 2007). The C-COSMOS survey has a complex tiling (see Fig. \\ref{overlay}) in comparison to other X-ray surveys, in which the overlapping areas of the single pointings are small and with similar PSFs (see e.g., the Extended Groth-Streep, AEGIS-X, Laird et al. 2009), or all the pointings are co-assial and nearly totally overlapping (see e.g., CDFS, Giacconi et al. 2001, Luo et al. 2008). In the C-COSMOS tiling, the pointings are strongly overlapping and not-coassial. While this ensures a very uniform sensitivity over most of the field, each source is observed with up to six different PSFs, requiring the development of an analysis procedure for data observed with this mixture of PSF. The procedures presented in this paper are aimed at optimizing (1) source detection, (2) localization, (3) photometry, and (4) survey sensitivity. We have made detailed comparisons between different detection algorithms and photometry tools, testing them extensively on simulated data. We furthermore validate our results by detailed inspections of each single source candidate. Our final analysis consists of a two main steps: \\begin{itemize} \\item [1] \\underline{a wavelet detection algorithm}, {\\it PWDetect} (Damiani et al. 1997) is first used to find source candidates. This algorithm is optimized to cleanly separate nearby sources, to detect point-like sources on top of extended emission and to give the most accurate positions. \\item [2] \\underline{A maximum likelihood PSF fitting algorithm} is then used to evaluate the source count rates and the probability that each source candidate is not a fluctuation of the background. We used the {\\it emldetect} algorithm (Cappelluti et al. 2007 and references therein). {\\it emldetect} works simultaneously with multiple overlapping pointings using PSFs appropriate to each one. This fitting method ensures accurate evaluation of the survey completeness and contamination, efficient deblending and good photometry for close pairs, which may be partly blended even at the {\\it Chandra} resolution. \\end{itemize} As a third step, we also performed \\underline{aperture photometry} for each candidate X-ray source using 50\\%, 90\\%, and 95\\% encircled count fractions, using the PSFs appropriate to each observation. The aperture photometry is also used to check the results. Aperture photometry is preferable in all cases where the systematic error introduced by PSF fitting are larger than the statistical errors, i.e., for bright sources (count rates $\\geq 1$ counts/ksec). The survey sensitivity is limited by both the net (i.e., including vignetting) exposure time, and by the actual PSF with which a given region of the area is observed. The latter issue is particularly relevant for the C-COSMOS tiling. We have developed an algorithm that evaluates the survey sensitivity at each position on C-COSMOS using a parameterization of the {\\it Chandra} ACIS-I PSF and taking into account the mixture of PSFs at each position. The resulting sensitivity maps have been compared and validated with extensive simulations. The paper is organized as following: in Sect. 2 we briefly present the C-COSMOS observations and data reduction; we describe the simulations in Sect. 3; how they were used to select the most efficient detection algorithm and the final source characterization procedure is described in Sect. 4; the completeness and reliability are shown in Sect. 5; in Sect. 6 we apply this procedure to the observed data; in Sect. 7 we present the calculation of survey sensitivity, the sky-coverage, and X-ray number counts using the simulated data. Finally, in Sect. 8 we compare C-COSMOS to a similar {\\it Chandra} survey, i.e., AEGIS-X, and in Sect. 9 we give our conclusion. ", "conclusions": "The complex tiling of C-COSMOS survey required the development of a tailored multistep procedure to fully exploit the data. Detailed simulations were used to test different detection (sliding cell and wavelet) and photometry (PSF fitting and aperture photometry) algorithms. In particular, we compared the results obtained using the SAS {\\it eboxdetect} and {\\it emldetect} tasks, used for the XMM-COSMOS survey (Cappelluti et al. 2007, 2009), with those obtained using the {\\it PWDetect} code (Damiani et al. 1997). Through these tests we selected a procedure consisting in first identifying source candidates using {\\it PWDetect}, and then performing accurate PSF fitting photometry and evaluating aperture photometry for each source candidate. In this way we obtained subarcsec source localizations and accurate photometry even for partly blended sources. We set a threshold for source detection to $P=2\\cdot 10^{-5}$, which implies a completeness of 87.5\\% and 68\\% for sources with at least 12 and 7 F band counts, respectively, and 3 to 5 spurious detections in the F band at the same count limits, respectively. We evaluated the survey sensitivity and the sky-coverage, through an analytical method, tuned using simulations. We then evaluated the $\\log N$ -- $\\log S$ of the detected sources in the simulations down to F, S, and H band flux limits of F$_{x}$$\\sim~2.3\\cdot 10^{-16}$, $\\sim~1.6\\cdot 10^{-15}$, and $\\sim~9.6\\cdot 10^{-16}$ erg s$^{-1}$ cm$^{-2}$, respectively. Finally we compared the C-COSMOS survey to the AEGIS-X survey, a {\\it Chandra} survey with similar sky-coverage and total exposure time, but using non overlapping ACIS-I pointings. We found that the complex tiling of C-COSMOS helps in obtaining a contiguous area with uniform sensitivity and somewhat higher source density. The overlap of several pointings with different PSF at the same position produces an effective source extraction region of $\\sim$~3 arcsec radius, i.e., significantly wider than the Chandra PSF at off-axis angles smaller than 5-6 arcmin. This produces a number of independent detection cells per unit area smaller than in a single ACIS-I pointing survey like AEGIS-X, which in turn implies a smaller number of spurious sources at each given detection threshold." }, "0910/0910.1614_arXiv.txt": { "abstract": "An analysis of archival mid-infrared (mid-IR) spectra of Seyfert galaxies from the \\textit{Spitzer Space Telescope} observations is presented. We characterize the nature of the mid-IR active nuclear continuum by subtracting a template starburst spectrum from the Seyfert spectra. The long wavelength part of the spectrum contains a strong contribution from the starburst-heated cool dust; this is used to effectively separate starburst-dominated Seyferts from those dominated by the active nuclear continuum. Within the latter category, the strength of the active nuclear continuum drops rapidly beyond $\\sim 20\\mum$. On average, type 2 Seyferts have weaker short-wavelength active nuclear continua as compared to type 1 Seyferts. Type 2 Seyferts can be divided into two types, those with strong poly-cyclic aromatic hydrocarbon (PAH) bands and those without. The latter type show polarized broad emission lines in their optical spectra. The PAH-dominated type 2 Seyferts and Seyfert 1.8/1.9s show very similar mid-IR spectra. However, after the subtraction of the starburst component, there is a striking similarity in the active nuclear continuum of all Seyfert optical types. PAH-dominated Seyfert 2s and Seyfert 1.8/1.9s tend to show weak active nuclear continua in general. A few type 2 Seyferts with weak/absent PAH bands show a bump in the spectrum between 15 and 20$\\mum$. We suggest that this bump is the peak of a warm ($\\sim$200$\\kel$) blackbody dust emission, which becomes clearly visible when the short-wavelength continuum is weaker. This warm blackbody emission is also observed in other Seyfert optical sub-types, suggesting a common origin in these active galactic nuclei. ", "introduction": "The complete mid-infrared (mid-IR) spectra of Seyfert galaxies \\citep[a class of active galactic nuclei (AGN):][]{1943ApJ....97...28S} have been available only in recent past \\citep[\\eg][]{2002A&A...393..821S,2005SSRv..119..355V,2005ApJ...633..706W,2007ApJ...671..124D}. Many {\\it Spitzer} spectra of nearby Seyfert galaxies show a strong contribution from star-forming features in the form of poly-cyclic aromatic hydrocarbon (PAH) bands \\citep[\\eg][]{2000A&A...357..839C,2006AJ....132..401B,2008ApJ...676..836T}. A number of these features, such as the 7.7$\\mum$ and the 17$\\mum$ PAH complex, are strongly blended with each other, complicating the estimation of the underlying active nuclear continuum. Further, the mid-IR opacity includes a strong contribution from the silicate bands at $10$ and $18\\mum$. Thus, estimating the intrinsic active nuclear continua in the mid-IR is a non-trivial task, with the primary hurdle being the subtraction of the starburst component. Previous studies have hinted that the continuum in the 1--8$\\mum$ range is non-stellar in origin and is likely a result of thermal emission from dust heated close to sublimation temperature by the optical/ultra-violet continuum from the central source \\citep{1986ApJ...308...59E,2001AJ....121.1369A,2004ApJ...614..122I,2008arXiv0807.4695M}. In a multi-wavelength photometric study of spectral energy distributions (SED) of predominantly radio-quiet Sloan Digital Sky Survey (SDSS) quasars, \\citet{2006ApJS..166..470R} and \\citet{2007ApJ...661...30G} noted that the 1--8$\\mum$ spectral index ($\\alpha_{\\nu}$) is strongly anti-correlated with infrared luminosity in type 1 quasars. The more luminous quasars have flatter 1--8$\\mum$ slopes. A linear correlation between the optical continuum luminosity and the infrared luminosity suggests that the observed bump around $\\sim2.2\\mum$ in the SED is driven by the dust re-emission. For example, the $2.2\\mum$ bump is clearly visible in the near-IR spectrum of Mrk~1239 \\citep{2006MNRAS.367L..57R}. Due to the presence of strong PAH bands in the 5--8$\\mum$ range in many Seyfert spectra, direct measurement of the active nuclear continuum in this region is only possible in sources with very weak or absent PAH features. An alternative is to subtract the starburst contribution using a template starburst spectrum and then study the residual continuum. We take this simple approach in this paper. In Section~2, we describe our archival sample and the data analysis techniques. We discuss the observed mid-IR spectra in Section~3. In Section~4, we use simple continuum diagnostics that allows us to classify Seyfert mid-IR spectra into PAH-dominated and AGN-dominated groups and understand continuum properties. In Section~5, we discuss the starburst contribution in Seyfert spectra and the resulting continuum shapes after subtraction of the starburst template spectrum. In Section~6, we summarize our results. ", "conclusions": "An analysis of archival {\\it Spitzer Space Telescope} mid-IR spectra of Seyfert galaxies is presented. We focus on understanding the intrinsic shape of the active nuclear continuum in the mid-IR region and how it relates to other properties of the source such as the 10$\\mum$ silicate optical depth. We assumed a template spectrum for the starburst component, and subtracted it from the Seyfert spectra to separate the active nuclear contribution from the circum-nuclear starburst contribution. Our primary conclusions from this study are as follows: \\begin{enumerate} \\item Seyfert spectra are classified effectively between AGN- and starburst-dominated categories based on the spectral indices, $\\alpha_{\\lambda}(5.5\\textrm{--}14.7\\mum)$ and $\\alpha_{\\lambda}(20\\textrm{--}30\\mum)$ (see Figure~\\ref{fig:continua}). Seyferts dominated by the AGN contribution have flatter spectra with $\\alpha_{\\lambda}(5.5\\textrm{--}14.7\\mum) \\sim -1.13$. The added starburst contribution from the host galaxy in the large {\\it Spitzer} aperture (see Table~\\ref{tab:objdata}) makes $\\alpha_{\\lambda}(20.0\\textrm{--}30.0\\mum)$ more positive or steeper. A key property that distinguishes AGN-dominated spectra is that the 20--30$\\mum$ continuum is flatter ($\\alpha_{\\lambda}\\sim -2$) than the very steep 20--30$\\mum$ continuum of starburst-dominated objects ($\\alpha_{\\lambda}\\sim 1$). The 20--30$\\mum$ continuum is likely formed from the Rayleigh--Jeans tail of the ``warm'' dust component. It is likely that there are multiple ``warm'' components with different temperatures. Further, Type 2 Seyferts with polarized broad emission lines in their optical spectra (type S1h) show $\\alpha_{\\lambda}(5.5\\textrm{--}14.7\\mum) \\sim - 0.49$ much different than average type 1 Seyferts that show $\\alpha_{\\lambda}(5.5\\textrm{--}14.7\\mum) \\sim -1.13$. Note the steeper and weak short-wavelength continuum in Figure~\\ref{fig:all_spec_nosb}, as compared to type 1 Seyferts. This is a direct evidence for presence of the dust torus that blocks our view of the hot dust closer in. \\item After starburst subtraction, Seyfert 1.8/1.9s and Seyfert 2s with strong PAH features in their spectra show similar active nuclear continuum as Seyfert 2s with weak/absent PAH features in their mid-IR spectra and polarized broad emission lines in their optical spectra (see Figure~\\ref{fig:all_spec_nosb}). This suggests presence of similar quantities and/or properties of dusty material around the central accretion disk in these type 2 sources. \\citet{2003ApJ...583..632T} had proposed existence of two types of Seyfert 2s: the HBLR and the non-HBLRs. We compared spectral indices in the 5--8$\\mum$ region after starburst subtraction and find that both the HBLR and non-HBLR show similar spectral indices (Figure~\\ref{fig:after-sb-sub}, bottom), suggesting similarity rather than differences between the two classifications. While, non-HBLRs tend to show stronger starburst contribution as compared to their AGN contribution, the converse (that starburst-dominated systems lack BLR signatures) is not necessarily true. The additional host galaxy contribution likely complicates the identification of BLR in these systems. As we show in Figure~\\ref{fig:continua}, simple continuum indices effectively separate AGN-dominated Seyferts from starburst-dominated Seyferts in the mid-IR. \\item \\citet{2007ApJ...671..124D} showed that a large part of the silicate absorption in some Seyfert galaxies comes from starburst-heated cold dust in the host galaxy rather than the dust torus. This conclusion was based on the fact that only highly inclined galaxies showed large silicate optical depths. Here, we put that result on a better statistical basis by presenting a correlation between the $b/a$ and the measured optical depth for this sample of 109 sources (see Figure~\\ref{fig:opt-depth-ba}). All objects with significant optical depth are highly inclined and are also classified as Seyfert 2s or Seyfert 1.8/1.9s. This confirms previous results by \\cite{1980AJ.....85..198K} and \\cite{1995ApJ...454...95M}, and highlights the importance of considering the host galaxy contribution in concealing AGN. \\item On average, Seyfert galaxies dominated by the AGN continuum tend to show weak silicate absorption ($\\tau_{9.7} \\lesssim 0.4$). The short wavelength continuum index ($\\alpha_{\\lambda}(5.5\\textrm{--}14.7\\mum)$) and the apparent silicate optical depth $\\tau_{9.7}$ (Figure~\\ref{fig:opt-depth}) may to be correlated in AGN-dominated objects. Seyfert optical types form a continuous sequence of increasing optical depth along this correlation from type 1s, type 1.8/1.9s, to type 2s with HBLRs. This validates the general inclination dependence inherent in AGN models noted before in the mid-IR by \\citet{2007ApJ...655L..77H}. But, as can be noted in Figure~\\ref{fig:opt-depth}, there is not a strict relationship between the strength of silicate features and the optical spectral type. \\item Figure~\\ref{fig:after-sb-sub} (top) shows that there are at least two types of dust distributions in the active nuclear region, one that generates the short-wavelength continua and another that generates the long-wavelength continua. The major difference between these two dust distributions is likely to be their mean temperatures. By subtracting the starburst and associated cool dust contribution, we have essentially removed the starburst component that contributes most to the variety in Seyfert spectra \\citep{2006AJ....132..401B}. In Figure~\\ref{fig:bump}, we demonstrate the existence of this ``warm'' component which dominates the long-wavelength continuum, by separating it from the hot dust component in Mrk 766. This simple template subtraction exercise provides proof that Seyfert spectra are primarily thermal in nature and composed of at least three thermal components with $T\\sim$ 1000, 200, and 60$\\kel$. The reported ``break'' in the spectrum at $\\sim$ 20$\\mum$ in type 1 Seyfert spectra is a result of this warm component being brighter at $\\sim 20 \\mum$ than the Rayleigh--Jeans tail of the hot component. The above-mentioned similarity of continua beyond $\\sim15\\mum$ in AGN-dominated sources is also due to this warm dust component being present in almost all observed AGN spectra. \\end{enumerate}" }, "0910/0910.3611_arXiv.txt": { "abstract": "In a dynamical-radiative model we recently developed to describe the physics of compact, GHz-Peaked-Spectrum (GPS) sources, the relativistic jets propagate across the inner, kpc-sized region of the host galaxy, while the electron population of the expanding lobes evolves and emits synchrotron and inverse-Compton (IC) radiation. Interstellar-medium gas clouds engulfed by the expanding lobes, and photoionized by the active nucleus, are responsible for the radio spectral turnover through free-free absorption (FFA) of the synchrotron photons. The model provides a description of the evolution of the spectral energy distribution (SED) of GPS sources with their expansion, predicting significant and complex high-energy emission, from the X-ray to the $\\gamma$-ray frequency domain. Here, we test this model with the broad-band SEDs of a sample of eleven X-ray emitting GPS galaxies with Compact-Symmetric-Object (CSO) morphology, and show that: (i) the shape of the radio continuum at frequencies lower than the spectral turnover is indeed well accounted for by the FFA mechanism; (ii) the observed X-ray spectra can be interpreted as non-thermal radiation produced via IC scattering of the local radiation fields off the lobe particles, providing a viable alternative to the thermal, accretion-disk dominated scenario. We also show that the relation between the hydrogen column densities derived from the X-ray (\\NHmath) and radio (\\NHImath) data of the sources is suggestive of a positive correlation, which, if confirmed by future observations, would provide further support to our scenario of high-energy emitting lobes. ", "introduction": "\\label{sec_introduction} The power released by active galactic nuclei (AGNs) is currently interpreted in terms of conversion of gravitational energy to radiative energy by accretion processes feeding the central supermassive black hole (BH) with environmental gas. The triggering, maintenance, and fading of the AGN activity, as well as the link of these processes with the physical conditions of the environment, from sub- to super-galactic scales, are still widely debated issues, and keep on stimulating a large variety of scientific investigations. In this context, radio galaxies are ideal laboratories, because they offer an edge-on view of both their nuclei and their relativistic jets, launched from the galactic center and reaching up to Mpc distances. The variety of powers, sizes, and morphologies displayed by radio galaxies not only provides pieces of evidence of different environmental physical conditions, but also samples subsequent stages of the source evolution. In particular, key sources for the investigation of the very first phases of the evolution of radio galaxies are the Gigahertz-Peaked-Spectrum (GPS) sources associated with galaxies and characterized by a Compact Symmetric Object (CSO) radio morphology. GPS sources \\citep[see][for a review]{odea1998} are a class of powerful radio sources ($P_{1.4\\, {\\rm GHz}}\\gtsim 10^{25}$ \\wattperhz) displaying convex radio spectra that turn over at frequencies of about 0.5--1 GHz; they make up a conspicuous fraction ($\\sim$10\\%) of the radio sources found in high-frequency radio surveys, and are optically identified with either galaxies or quasars; their radio sizes are smaller than about 1 kpc, and their radio morphologies reveal either core-jet structures or mini-lobes embedding terminal hotspots and possibly straddling a central core. CSOs \\citep{wilkinson1994,readhead1996} have instead emerged in VLBI surveys as a class of radio sources with compact ($\\ltsim$500 pc) and symmetric radio structures, making them resemble ``classical double'' radio galaxies in miniature. Whereas there is no debate on the true compactness of CSOs, the compactness of a fraction of the core-jet GPS sources might be the result of foreshortening of extended radio sources roughly aligned with the line of sight; especially (although not exclusively) in these cases, the GPS itself might be a transient spectral state: flux-density variability and polarization studies at different radio frequencies are thus instrumental to the selection of bona-fide samples of GPS sources \\citep{tinti2005,torniainen2005,torniainen2007,orienti2008a}. The overlap between the GPS-source and the CSO classes is however significant: GPS sources associated with galaxies are most likely to display a CSO morphology, and most CSOs exhibit a GPS. This overlap is explained in terms of synchrotron radio spectra dominated by the emission of the mini-lobes, and suffering from absorption effects causing the turnover about 1 GHz \\citep{snellen2000}. GPS/CSOs {associated with} galaxies are thus high-confidence candidates for truly compact sources. Although compact GPS sources were proposed to be young objects soon after their discovery \\citep[e.g.,][]{shklovsky1965}, a widely discussed alternative to the youth scenario was the frustration scenario, ascribing the small source size to confinement by a particularly dense interstellar medium (ISM) preventing the jet expansion \\citep{vanbreugel1984}. Because the required ISM confining densities are not confirmed by recent studies \\citep[e.g.,][]{morganti2008}, the confinement scenario is considered less likely, although it might still apply to selected objects \\citep[e.g.,][]{garciaburillo2007}. Much observational evidence has instead accumulated in favour of the youth scenario, the most compelling measurement being the detection, in several GPS/CSO galaxies, of hotspot advance velocities about 0.1--0.2$c$ \\citep{owsianikconway1998,owsianik1998,tschager2000,taylor2000,gugliucci2005}, which indicate kinematical source ages not higher than a few $10^3$ years, in good agreement with spectral ageing estimates \\citep{murgia1999}. Further evidence of youth comes from the underluminosity of GPS/CSO optical narrow-emission lines, suggesting that the Str\\\"omgren sphere of the recently-triggered AGN is still in an expansion phase in these sources, and we are thus witnessing the birth of their Narrow-Line Region \\citep[NLR;][]{vink2006}. The youth scenario is part of a wider evolutionary scenario \\citep{fanti1995,snellen2000}, according to which GPS/CSOs would first evolve into symmetric Compact-Steep-Spectrum (CSS) sources, equally powerful but with sizes of about 1--15 kpc, and then further expand outside the host galaxy to become large-scale, powerful radio galaxies. The evolutionary link between GPS and CSS sources is supported by their membership to the anticorrelation between linear size and peak frequency \\citep[$LS \\propto \\nu_{\\rm p}^{-0.65}$;][]{odea1997}. However, it is not clear yet whether all GPS/CSOs will eventually become large-scale radio sources, and whether the evolved sources will display a Fanaroff-Riley type I (FRI) or type II (FRII) morphology, because of the frequency's observational limits in the exploration of this relation for super-galactic sized sources, as well as because of the biases in the surveys from which the distribution of the radio sources in the power-linear size ($P-LS$) diagram is drawn \\citep[e.g.,][and references therein]{snellen2000}. The first systematic studies of the host galaxies of GPS sources, conducted in the optical and near-infrared (NIR) bands \\citep{snellen1996,odea1996,devries1998a,devries1998b,devries2000} showed that they are, as CSS sources, characterized by luminous masses, brightness profiles, and optical-NIR colors more typical of passively- or non-evolving giant elliptical galaxies, than either spirals, small ellipticals, brightest cluster galaxies (BCGs), or central-dominant (cD) galaxies at similar redshifts. This seems to be consistent with what was found for a sample of FRIIs rather than with the properties of FRIs, which are preferentially hosted by cD galaxies \\citep{owen1989,owen1991,zirbel1996}, and suggest that GPS sources are more likely to evolve, through the CSS phase, in FRIIs rather than in FRIs \\citep{devries2003}. However, the morphologies of the majority \\citep[$\\sim$60\\%;][]{odea1996} of the hosts of GPS sources show signs of recent mergers and/or interactions. This evidence, apparently at odds with the passively-evolving scenario, was first interpreted as an indication that GPS sources might be associated with the first of a sequence of mergers \\citep{devries2003}, possibly leading to the formation of a BCG or a cD only over time scales much longer than the lifetime of the radio source. In subsequent investigations, however, the presence of young stellar population in these giant ellipticals was revealed by means of stellar-population synthesis models applied to photometric \\citep{devries2007} and spectrophotometric data \\citep{holt2009}. Furthermore, \\citet{devries2007} showed the similarity, out to $z \\simeq 0.55$, between the luminosities of the GPS-source hosts and of the Luminous Red Galaxies (LRGs), which are thought to represent the most massive early-type galaxies, and are often associated with BCGs. Therefore, the properties of the hosts do not provide unambiguous indications on the subsequent GPS/CSO evolutionary stages. Despite these results, population studies highlight the existence of far too many compact sources compared to the number of powerful large-scale objects \\citep{odea1997}. Beside the possibility, not easy to justify, that these sources undergo a short-lived activity \\citep{readhead1996}, two scenarios prove to be particularly attractive to explain this statistical evidence. In the first scenario, the jet activity is intermittent on time scales comparable to the age of the small sources \\citep[e.g.,][]{reynolds1997}: this scenario, supported by the observations of double-double radio galaxies \\citep{lara1999,schoenmakers2000,kaiser2000}, and by the detection of candidates for dying compact sources \\citep{giroletti2005,parma2007}, finds a natural explanation in the framework of accretion-disk instabilities \\citep{czerny2009}. An accretion disk operating above a given threshold accretion rate ($\\dot m_* \\simeq 0.025\\, \\dot m_{Edd}$) is expected to experience instabilities driven by the radiation pressure, causing the source undergo alternate phases of high and low activity; for moderate accretion rates, the duration of the active phases is short enough ($\\sim 10^3-10^4$ years) to make the compact source unable to grow to super-galactic sizes, whereas accretion rates close to the Eddington limit are required for the development of large-scale sources \\citep{czerny2009}. In the second scenario, many sources undergo processes of jet-flow turbulent disruption before their lobes can grow to large sizes, either disappearing from the radio sky \\citep{alexander2000} or developing an FRI-type morphology \\citep{kaiser2007}, in both cases experiencing a drop in luminosity. The recent finding that FRIs smaller than $\\sim$40 kpc were almost absent in a sample of radio galaxies optically identified in the SDSS \\citep{best2009} seems to suggest that all radio sources begin their life with collimated jets, but less powerful jets are more easily disrupted, as the source grows, in denser environments, leading to the production of FRI sources. Further evidence supporting this view might derive from the confirmation that the large fraction of low-power, compact objects with non-FRI morphology discovered within a sample of FRI candidates at $13$ s) GRBs with hardness above and below the median Swift $E_{\\rm pk,obs}= 100$ keV and also short duration ($T_{90}<3$ s) GRBs. The rate of long-duration hard GRBs is turning over at low flux levels, while the rate of long-duration soft GRBs rises more strongly. This is a gradual effect in $E_{\\rm pk,obs}$. Although the logN--logS slope for long-duration GRBs does not appear to be duration dependent, the Swift short-duration GRB population is strongly rising in number to low flux levels, showing no significant sign of a turn-over. The curves expected without a cutoff --- derived in Section \\ref{sec:fitting} --- due to the detector are plotted as dotted lines. The dashed red curve (barely visible) at the left of the short-duration curve accounts for the detector threshold following the non-parametric prescription of \\citet{pet96}. } \\label{fig:logN-logS} \\end{figure} \\subsection{Prior Rate and Luminosity Function Estimates} \\label{sec:prior} There is a rich literature describing optimal ways of counting GRBs to determine their distance and intrinsic flux. In the pre-afterglow era, counting focused on the observed flux distributions. The number of events $N$ with observed flux greater than $S$ --- the so called ``logN--logS'' curve --- showed early evidence (over many decades in $S$) for slope $S^{-3/2}$ expected for a homogeneous, isotropic, and static source population in a Euclidean universe \\citep[HISE; e.g.,][]{hurley91,hs90}. A powerful statistic for examining the source counts is $V/V_{\\rm max} = (C/C_{\\rm min})^{-3/2}$ \\citep{schmidt68}, a measure of the volume probed by a source detected with $C$ counts relative to a possible minimum number of observable counts $C_{\\rm min}$. The expectation is that $\\langle V/V_{\\rm max} \\rangle = 0.5$ for HISE. The BATSE experiment provided the first strong evidence from a single experiment \\citep{meegan92} --- a deficit of low $S$ GRBS and $\\langle V/V_{\\rm max} \\rangle <0.5$ --- for a departure from homogeneity, while the spatial counts showed clearly an isotropic population. To study whether these modest departures imply that GRBs are very local (a Galactic halo population) or cosmological required examination beyond the first moment $\\langle V/V_{\\rm max} \\rangle$ in the $V$ (or $C$) distribution \\citep[e.g.,][]{band92,ht93,pet93}. The first GRB redshifts \\citep[e.g.,][]{metzger97} defined a cosmological origin and a vast energy release. Connecting the small number of GRBs with $z$ to the large population of GRBs without $z$ required, in general, careful modelling of and strong assumptions for the intrinsic luminosity and number density distributions in order to reproduce the observed flux data \\citep[e.g.,][]{piran92, piran99, cp95, fb95, lw95, lw98, hh97, schmidt99, schmidt01, sb01, guetta05}. Exceptions to the parametric approach were studies utilizing luminosity criteria (i.e., possible correlations of observables with luminosity) to derive ``pseudo-redshifts'' for the full GRB sample \\citep[e.g.,][]{norris00,fr00,schaf01,lloyd02,mur03,yon04,firm04,kocevski06,schmidt09}. These studies generally found a rising GRB rate to $z \\lessim 2$, similar to the cosmic star formation rate \\citep[e.g.,][]{madau96}, but potentially continuing to remain flat or even rising to $z\\sim 12$. Notably, some of these works \\citep[e.g.,][]{lloyd02,yon04}, using hazard statistics \\citep{lb71,ep92,pet93,mp99}, also found evidence for potential strong luminosity evolution, parameterized as $L\\propto (1+z)^a$, with $a$ in the range 1.5--2.5. The luminosity function itself appears to generally be characterized well as broken powerlaw, with a break at $L\\sim 10^{51-52}$ erg s$^{-1}$ and a flat or slowly rising slope to low-energies, strongly dependent upon the instrumental detection model. The connection between GRBs and the deaths of massive stars (now firmly established, e.g., \\citet{stanek03,hjorth03}; see \\citet{woosbloom06} for a review) sped progress by motivating an assumption that the GRB rate follows star formation \\citep[e.g.,][]{wijers98,lr00,pm01,cs02,bloom03,gor04,nat05}. Very recently, thanks to Swift and the impressive efforts of ground-based observers, a growing sample of GRBs with spectroscopic redshifts has allowed for direct tabulation of GRB intrinsic luminosities \\citep[e.g.,][]{kocevskibutler08} and redshifts \\citep[e.g.,][]{jak05,jak06}. The large number of redshifts has also enabled a detailed comparison of the intrinsic GRB rate to the cosmic star formation rate \\citep[SFR, e.g.,][]{daigne06,salv07,kistler08,kistler09,salv09a,salv09b}. Perhaps the most intriguing, shared feature of these studies is a strong indication of evolution in the GRB population. Above $z\\approx 2$ and possibly extending to $z\\approx 8$, the Swit GRB rate is increasing far faster than star formation \\citep[e.g.,][]{kistler08,kistler09}, and it is not clear to what extent this is due to GRBs in the early universe being bright \\citep[i.e., luminosity evolution, preferred by][]{salv07,salv09a,salv09b,pet09} or to an increase in the overall number of GRBs at intermediate and high-$z$ relative to the SFR. As we discuss below in Section \\ref{sec:rho}, rigorous treatment of the largest available Swift dataset (Section \\ref{sec:redux}) allows for a firm conclusion in favor of rate evolution and not luminosity evolution, and we suggest plausible explanations. To draw this conclusion and to study GRB rates as a function of intrinsic hardness, flux, and duration (Sections \\ref{sec:fitting} \\& \\ref{sec:energetics}) as well as $z$, we require a detailed model for the Swift satellite detection limit (Section \\ref{sec:limit}). ", "conclusions": "" }, "0910/0910.3946_arXiv.txt": { "abstract": "We present deep CFHT/MegaCam photometry of the ultra-faint Milky Way satellite galaxies Coma Berenices (ComBer) and Ursa Major II (UMa\\,II). These data extend to $r\\sim25$, corresponding to three magnitudes below the main sequence turn-offs in these galaxies. We robustly calculate a total luminosity of $M_{V}=-3.8\\pm0.6$ for ComBer and $M_{V}=-3.9\\pm0.5$ for UMa\\,II, in agreement with previous results. ComBer shows a fairly regular morphology with no signs of active tidal stripping down to a surface brightness limit of $32.4$ mag arcsec$^{-2}$. Using a maximum likelihood analysis, we calculate the half-light radius of ComBer to be $r_{\\rm half}=74\\pm4$\\,pc ($5.8\\pm0.3\\arcmin$) and its ellipticity $\\epsilon=0.36\\pm0.04$. In contrast, UMa\\,II shows signs of on-going disruption. We map its morphology down to $\\mu_{V}=32.6$ mag arcsec$^{-2}$ and found that UMa\\,II is larger than previously determined, extending at least $\\sim700$\\,pc ($1.2^\\circ$ on the sky) and it is also quite elongated with an ellipticity of $\\epsilon=0.50\\pm0.2$. However, our estimate for the half-light radius, $123\\pm3$\\,pc ($14.1\\pm0.3\\arcmin$) is similar to previous results. We discuss the implications of these findings in the context of potential indirect dark matter detections and galaxy formation. We conclude that while ComBer appears to be a stable dwarf galaxy, UMa\\,II shows signs of on-going tidal interaction. ", "introduction": "For decades, only about a dozen dwarf galaxies were known to orbit the Milky Way. The majority of these systems corresponded to dwarf spheroidal (dSph) galaxies, the least luminous, but, by number, the dominant galaxy type in the present-day universe. However, over the last five years, and thanks to the advent of the Sloan Digital Sky Survey (SDSS; \\citealt{York2000}) the field of dwarf galaxies in the Milky Way has been revolutionized. To date, fourteen new systems have been detected as slight overdensities in star count maps using the SDSS data (\\citealt{Willman2005a,Willman2005b}; \\citealt{Belokurov2006,Belokurov2007a,Belokurov2008,Belokurov2009}; \\citealt{Zucker2006a,Zucker2006b}; \\citealt{Sakamoto2006}; \\citealt{Irwin2007}; \\citealt{Walsh2009}). These recent discoveries have revealed a previously unknown population of ``ultra-faint'' systems which have extreme low luminosities, in some cases as low as $L_{V}\\sim300$\\,L$_{\\sun}$ \\citep{Martin2008}, and in average comparable (or lower) to those of Galactic globular clusters. However, spectroscopic surveys of the majority of these systems reveal kinematics and metallicities in line with those of dwarf galaxies (\\citealt{Kleyna2005}; \\citealt{Munoz2006}; \\citealt{SimonGeha2007}; \\citealt{Kirby2008}; \\citealt{Koch2009}; \\citealt{Geha2009}). Dynamical mass estimates of the ultra-faint galaxies based on line-of-sight radial velocities indicate that these galaxies are extremely dark matter dominated, with central mass-to-light ratios (M/L) as high as $1,000$ in solar units. These systems are thus good laboratories to constrain cosmological models \\citep[e.g~the 'missing satellites problem';][]{Kauffmann1993,Moore1999,Klypin1999,SimonGeha2007} and to study the properties of dark matter (\\citealt{Strigari2008}; \\citealt{Kuhlen2008}; \\citealt{Geha2009}). However, these applications hinge critically on the assumption that the masses and density distributions in these systems are accurately known. Current mass estimates for the ultra-faint dwarfs are based on the assumption that the dynamical state of these systems have not been significantly affected by Galactic tides, and therefore that they are near dynamical equilibrium. There is circumstantial evidence for past tidal disturbance in a number of these satellites based on morphological studies. \\citet{Coleman2007} for instance, found the Hercules dSph to be highly elongated, with a major to minor axis ratio of 6:1, larger than for any of the ``classical'' dSphs, and argued for a tidal origin of this elongation. \\citet{Belokurov2007a} reported fairly distorted morphologies for several of the new ultra-faint dwarfs, although \\citet{Martin2008} showed that results based on low number of star may not be statistically significant because they suffer from shot noise. The strongest evidence for tidal interaction perhaps is for Ursa Major II, which appears to be broken into several clumps, lies close to the great circle that includes the Orphan Stream (\\citealt{Zucker2006b}; \\citealt{Fellhauer2007}) and shows a velocity gradient along its major axis \\citep{SimonGeha2007}. If dSphs galaxies are currently undergoing tidal stripping, or have in the past, then kinematical samples are expected to be ``contaminated'' with unbound or marginally bound stars. This will impact subsequent dynamical modeling, often resulting in overestimated mass contents (\\citealt{Klimentowski2007}; \\citealt{Lokas2009}). The degree of such contamination is a function of both the projection of the orbit along the line-of-sight and the strength of the tidal interaction. Therefore, even though tides do not affect appreciably the inner kinematics of dSphs until the latest stages of tidal disruption (e.g., \\citealt{OLA95}; \\citealt{Johnston1999}; \\citealt{Munoz2008}; \\citealt{Penarrubia2008}), studies aimed at identifying the presence of tidal debris can help elucidate the dynamical state of these dwarf galaxies. Obvious tidal features around dSphs would indicate the presence of unbound stars and their effects on the derived masses would need to be investigated. Alternatively, a lack of clear detections in the plane of the sky would narrow the possibilities that kinematical samples suffer from contamination, although it would not automatically imply that an object has not been tidally affected. Tidal features could still exist but be aligned preferentially along the line-of-sight and thus be hard to detect. Ultra-faint dwarf galaxies are useful probes of galaxy formation on the smallest scales (\\citealt{Madau2008}; \\citealt{Ricotti2008}). One of the outstanding questions related to their discovery is whether these galaxies formed intrinsically with such low luminosities, or whether they were born as brighter objects and attained their current luminosities through tidal mass loss. Current metallicity measurements (e.g., \\citealt{Kirby2008}) support the former scenario. They show that the ultra-faint dwarfs are also the most metal poor of the dSphs, following the luminosity-metallicity trend found for the classical dSphs\\footnote{\\citet{SimonGeha2007}, using the same spectroscopic data but a different technique, reported systematically higher metallicites for the ultra-faint dwarfs than those of \\citet{Kirby2008}. The discrepancy is explained by the fact that the former study used the \\citet{Rutledge1997} method that relies on the linear relationship between Ca II triplet equivalent widths and [Fe/H], a technique now known to fail at very low metallicities.}. On the other hand, if tidal features around these objects are firmly detected, it would clearly support the latter hypothesis. Given the importance of the questions at hand, it is essential that we investigate the dynamical state of these satellites. Due to the extreme low luminosity of the ultra-faint dwarfs, and therefore low number of brighter stars available for spectroscopic studies, deep imaging is currently the only way to efficiently detect the presence of faint morphological features in the outskirts of these systems. We expect any tidal debris to be of very low surface brightness (\\citealt{Bullock2005}) so that detecting it via integrated light is virtually impossible. However, these galaxies are sufficiently nearby to resolve individual stars. Thus, using matched-filter techniques it is possible to detect arbitrarily low surface brightness features ($\\sim 35$ mag arcsec$^{-2}$, \\citealt{Rockosi2002}; \\citealt{Grillmair2006}). In this article we present the results of a deep, wide-field photometric survey of the Coma Berenices (ComBer) and Ursa Major II (UMa\\,II) dSphs, claimed to be two of the most dark matter dominated dSphs \\citep{Strigari2008}, carried out with the MegaCam imager on the Canada-France-Hawaii Telescope. In \\S2, we present details about the observations and data reduction, as well as of artificial star tests carried out to determine our completeness levels and photometric uncertainties. In \\S3, we recalculate the structural parameters of these systems using a maximum likelihood analysis similar to that of \\citet{Martin2008}. In \\S4 we present the results of our morphological study. We discuss the statistical significance of our density contour maps and the robustness of our result to the variation of different parameters. We show that ComBer looks fairly regular in shape and find no signs of tidal debris down to a surface brightness limit of 32.4 mag arcsec$^{-2}$, whereas UMa\\,II shows a significantly elongated and distorted shape, likely the result of tidal interaction with the Milky Way. Our discussion and conclusions are presented in \\S5 and \\S6, respectively. ", "conclusions": "We have carried out a deep, wide-field photometric survey of the Coma Berenices and Ursa Major II dwarf spheroidal galaxies using the MegaCam Imager on CFHT, reaching down to $r\\sim25$ mag, more than three magnitudes below the main sequence turn-offs of these Galactic satellites. This increases roughly by an order of magnitude, with respect to the original SDSS photometry, the number of stars that belong to the respective galaxies and that are available determination of their structural properties and for morphological studies. Our results can be summarized as follows: 1. We used a maximum likelihood analysis similar to the one used by \\citet{Martin2008} and \\citet{Sand2009} to calculate structural parameters for three different density profiles: King, Plummer and exponential. We find characteristic sizes of $r_{\\rm half}=74\\pm4$\\,pc ($5.8\\pm0.3\\arcmin$) and $123\\pm3$\\,pc ($14.1\\pm0.3\\arcmin$) for ComBer and UMa\\,II respectively (from the exponential profile). Our results provide much tighter constraints on these structural parameters than possible with previous datasets, but are consistent with earlier determinations using SDSS photometry. 2. We have re-calculated the total luminosities for both systems and find, for ComBer $M_{V}=-3.8\\pm0.6$ and for UMa\\,II $M_{V}=-3.9\\pm0.5$ (for a choice of Salpeter IMF), in very good agreement with previous results, confirming that ComBer and UMa\\,II are among the faintest of the known dwarf satellites of the Milky Way. We have also used a \\citet{Chabrier2001} IMF but the results remain virtually unchanged. 3. We have found that ComBer shows a fairly regular morphology with no clear detection of potential stripped material down to a surface brightness of $32.4$ mag arcsec$^{-2}$. Additionally, its number density profile is reasonably well matched by a choice of either King, Plummer or exponential profile. We thus conclude that ComBer is likely a stable dwarf galaxy which would make it one of the most dark matter dominated of the dSph systems. 4. We have also studied the morphology of UMa\\,II and find that, unlike ComBer, it shows signs of being significantly disrupted. UMa\\,II is larger than previously determined, extending at least $\\sim700$\\,pc ($1.2^\\circ$ on the sky) and it is also quite elongated. Its density profile and overall shape resemble a structure possibly in the latest stages of tidal destruction. Furthermore, its number density profile is not well matched by neither of the three profiles we tried and it is much better described by two power laws, further supporting a tidal scenario. 5. The overall 2D surface density distributions of both systems are not affected by shot-noise and are therefore robust. We find no evidence for isolated tidal debris beyond the main bodies of ComBer and UMa\\,II to our surface brightness limits of $32.4$ and $34.8$ mag arcsec$^{-2}$, respectively. Deep, wide-field imaging of the recently discovered ultra-faint galaxies currently lags behind spectroscopic observations of these objects. We show in this paper that high quality, deep photometry is an equally important tool in studying the dynamical state of the ultra-faint dwarfs. These data can also be used to constrain the star formation histories of ComBer and UMa\\,II which we will explore in a future contribution. We acknowledge David Sand, Josh Simon and Gail Gutowski for useful discussions and Peter Stetson for graciously providing copies of DAOPHOT and ALLFRAME. M.G and B.W. acknowledge support from the National Science Foundation under award number AST-0908752." }, "0910/0910.3207_arXiv.txt": { "abstract": "We lay out the framework to numerically study nonlinear structure formation in the context of scalar-field-coupled cold dark matter models ($\\varphi$CDM models) where the scalar field $\\varphi$ serves as dynamical dark energy. Adopting parameters for the scalar field which leave negligible effects on the CMB spectrum, we generate the initial conditions for our $N$-body simulations. The simulations follow the spatial distributions of dark matter and the scalar field, solving their equations of motion using a multilevel adaptive grid technique. We show that the spatial configuration of the scalar field depends sensitively on the local density field. The $\\varphi$CDM model differs from standard $\\Lambda$CDM at small scales with observable modifications of, e.g., the mass function of halos as well as the matter power spectrum. Nevertheless, the predictions of both models for the Hubble expansion and the CMB spectrum are virtually indistinguishable. Hence, galaxy cluster counts and weak lensing observations, which probe structure formation at small scales, are needed to falsify this class of models. ", "introduction": "The origin and nature of dark energy (Copeland~et~al.~2006) is one of the most difficult challenges facing physicists and cosmologists at the present time. Among all the proposed models to tackle this problem, the introduction of a scalar field is perhaps the most popular. The scalar field, denoted by $\\varphi$, should have no coupling to normal matter to be consistent with stringent constraints from experiments (Will 2006, and references therein), but could couple to the dark matter, therefore producing a fifth force between dark matter particles. This idea has gained a lot of interest in recent years because dark matter physics are unknown, and such a coupling could alleviate the coincidence problem of dark energy (e.g., Amendola 2000; Chiba 2001; Chimento et al. 2003). Furthermore, it is commonly predicted by low energy effective theories derived from a more fundamental theory. A specific and interesting possibility is the chameleon mechanism (Khoury \\& Weltman 2004; Mota \\& Shaw 2006), by virtue of which the scalar field acquires a large mass in high density regions and thus the fifth force becomes undetectable on short ranges, thus also evading constraints from the large-scale cosmic microwave background (CMB). Indeed, at the linear perturbation level, there have been a lot of studies about the coupled scalar field and $f(R)$ gravity models (e.~g., Li \\& Barrow 2007; Hu \\& Sawicki 2007). Nevertheless, little is known about these models on nonlinear scales. It is well known that the matter distribution at late times, i.e. $z\\lesssim 2$ for cluster scales, evolves in a nonlinear way, making the behavior of the scalar field more complex and the linear analysis insufficient to produce accurate results that can be confronted with observations. For the latter purpose, the best way forward is to perform full $N$-body simulations (Bertschinger 1998) to evolve the individual particles step by step. $N$-body simulations including scalar fields and related models have been performed before (Linder \\& Jenkins 2003; Mainini~et~al. 2003; Macci\\`o~et.~al. 2004; Springel \\& Farrar 2007; Kesden \\& Kamionkowski 2006a, 2006b; Farrar \\& Rosen 2007; Baldi~et~al. 2008; Oyaizu 2008; Keselman~et~al. 2009; Li \\& Zhao 2009). For example, in the work of Macci\\`o~et~al.~(2004) the simulations included several effects due to the coupling between dark energy and dark matter (e.g. modified gravitational constant, an extra dragging term in Newton's equations and time variable dark matter particle masses), but did not consider a spatial variation of the dark energy scalar field. The more complete simulation of the scalar field by Li \\& Zhao (2009) shows that this approximation is only good for a limited choice of parameters and the scalar field potential. Here we extend the work of Li \\& Zhao (2009). This paper is organized as follows: In \\S~\\ref{sect:description}, we shall briefly review the general equations of motion for the coupled scalar field model introduced in Li \\& Zhao (2009), and present our specific choices of the coupling function and the scalar field potential. In \\S~\\ref{sect:non-linear}, we describe the formulae and the algorithm of the $N$-body simulation, analyze the results of our coupled scalar field $N$-body simulations, compare it with that of the standard $\\Lambda$CDM model, and explain the physical origin of the new features. Finally, we conclude and discuss observational implications in \\S~\\ref{sect:disc}. ", "conclusions": "\\label{sect:disc} We have presented a general framework to study nonlinear structure formation in coupled scalar field models, in particular the models of Li \\& Zhao (2009). While these models are virtually indistinguishable from $\\Lambda$CDM on both very large and very small scales, intermediate scales at low redshift ($z\\lesssim 1$) relevant for galaxy clusters ($\\sim 10^{2}-10^{3}$kpc) open a new window to test and constrain the interesting part of the parameter space. On these scales, the matter power spectrum is significantly increased compared to that of $\\Lambda$CDM. Observationally, this would most likely appear as a change of $\\sigma_{8}$ on the order of $15$-$20$\\% for models with $\\gamma=0.5-1$ and $\\mu=10^{-6}$ (see Fig.~2). Any variation of $\\sigma_8$ seems to be lower than 30\\% for current lensing measurements such as the CFHT Legacy Survey (e.g., Hoekstra et al. 2006; Fig.~11 of Fu et al. 2008) over a rather limited range; however, future surveys, such as the Kilo-Degree Survey (KIDS), will be able to measure the scale dependence within the range $k=0.1-10h$Mpc$^{-1}$, where the deviation of the models from $\\Lambda$CDM is maximal." }, "0910/0910.4611_arXiv.txt": { "abstract": "We present an improved analysis of halo substructure traced by RR Lyrae stars in the SDSS stripe 82 region. With the addition of SDSS-II data, a revised selection method based on new $ugriz$ light curve templates results in a sample of 483 RR Lyrae stars that is essentially free of contamination. The main result from our first study persists: the spatial distribution of halo stars at galactocentric distances 5--100 kpc is highly inhomogeneous. At least 20\\% of halo stars within 30 kpc from the Galactic center can be statistically associated with substructure. We present strong direct evidence, based on both RR Lyrae stars and main sequence stars, that the halo stellar number density profile significantly steepens beyond a Galactocentric distance of $\\sim$30 kpc, and a larger fraction of the stars are associated with substructure. By using a novel method that simultaneously combines data for RR Lyrae and main sequence stars, and using photometric metallicity estimates for main sequence stars derived from deep co-added $u$-band data, we measure the metallicity of the Sagittarius dSph tidal stream (trailing arm) towards R.A.$\\sim2^h-3^h$ and $Dec\\sim0^\\circ$ to be 0.3 dex higher ($[Fe/H]=-1.2$) than that of surrounding halo field stars. Together with a similar result for another major halo substructure, the Monoceros stream, these results support theoretical predictions that an early forming, smooth inner halo, is metal poor compared to high surface brightness material that have been accreted onto a later-forming outer halo. The mean metallicity of stars in the outer halo that are not associated with detectable clumps may still be more metal-poor than the bulk of inner-halo stars, as has been argued from other data sets. ", "introduction": "} Studies of the Galactic halo can help constrain the formation history of the Milky Way and the galaxy formation process in general. For example, within the framework of hierarchical galaxy formation \\citep{fbh02}, the spheroidal component of the luminous matter should reveal substructures such as tidal tails and streams \\citep{jhb96,hw99,bkw01,har01}. The number of these substructures, due to mergers and accretion over the Galaxy's lifetime, may provide a crucial test for proposed solutions to the ``missing satellite'' problem \\citep{bkw01}. Substructures are expected to be ubiquitous in the outer halo (galactocentric distance $>15-20$ kpc), where the dynamical timescales are sufficiently long for them to remain spatially coherent \\citep{jhb96,may02}, and indeed many have been discovered \\citep[e.g.,][]{iba97,yan00,ive00,bel06,gri06,vz06,bel07a,new07,jur08}. Various luminous tracers, such as main sequence turn-off stars, RR Lyrae variables, or red giants are used to detect halo substructures, and of them, RR Lyrae stars have proven to be especially useful. RR Lyrae stars represent a fair sample of the old halo population \\citep{smi95}. They are nearly standard candles ($\\langle M_V \\rangle = 0.59 \\pm 0.03$ at $[Fe/H]=-1.5$, \\citealt{cc03}), are sufficiently bright to be detected at large distances ($5-120$ kpc for $14 < V < 21$), and are sufficiently numerous to trace the halo substructure with good spatial resolution ($\\sim1.5$ kpc at 30 kpc). Fairly complete and relatively clean samples of RR Lyrae stars can be selected using single-epoch colors \\citep{ive05}, and if multi-epoch data exist, using variability (\\citealt{ive00}; QUEST, \\citealt{viv01}; \\citealt{ses07}, \\citealt{dle08}; SEKBO, \\citealt{kel08}; LONEOS-I, \\citealt{mic08}). A useful comparison of recent RR Lyrae surveys in terms of their sky coverage, distance limits, and sample size is presented by \\citet{kel08} (see their Table 1). Compared to other surveys, the so-called stripe 82 survey (an equatorial strip defined by declination limits of $\\pm$1.27$^\\circ$ and extending from R.A.$\\sim$20$^h$ to R.A.$\\sim$4$^h$) based on Sloan Digital Sky Survey data (SDSS; \\citealt{yor00}) is the deepest one yet obtained (probing distances to $\\sim110$ kpc), and the only survey with available 5-band photometry. In an initial study based on SDSS stripe 82 data, we discovered a complex spatial distribution of halo RR Lyrae stars at distances ranging from 5 kpc to $\\sim$100 kpc \\citep[hereafter S07]{ses07}. The candidate RR Lyrae sample was selected using colors and low-order light curve statistics. In this paper, we present a full light curve analysis enabled by the addition of new stripe 82 observations. The resulting sample of RR Lyrae stars is highly complete and essentially free of contamination. The extended observations now allow the determination of flux-averaged magnitudes, thus providing better luminosity and distance estimates. This reduces the uncertainty in the spatial distribution of RR Lyrae stars, and increases the overall accuracy of computed density maps. While this paper was in preparation, another study of RR Lyrae stars in stripe 82, based on essentially the same data set, was announced \\citep{wat09}. In addition to analysis that overlaps with their work, here we also discuss \\begin{itemize} \\item An RR Lyrae template construction for the SDSS bandpasses, and analysis of the template behavior \\item Template fitting and visual inspection of best-fits for single-band and color light curves, which results in an exceedingly clean final sample \\item Extended color range for the initial selection, which results in about 20\\% more stars than in the sample analyzed by Watkins et al. \\item A step-by-step analysis of the sample incompleteness at the faint end \\item A new method for constraining the metallicity of spatially coherent structures that are detected using main sequence and RR Lyrae stars \\item A comparison of the observed RR Lyrae spatial distribution with predictions based on an oblate halo model constrained by SDSS observations of main sequence stars. \\end{itemize} We emphasize, however, that both studies obtain consistent main results: strong substructure in the spatial distribution of halo RR Lyrae stars and evidence for a steepening of the best-fit power-law description beyond a galactocentric radius of $\\sim$30 kpc. Our data set and the initial selection of candidate RR Lyrae stars are described in \\S~\\ref{data}, while the construction of light curve templates is presented in \\S~\\ref{ugriz_templates}. Various properties of the resulting RR Lyrae sample, with emphasis on the spatial distribution, are analyzed in \\S~\\ref{sec:spatial}. In \\S~\\ref{sec:MS} we compare the spatial distribution of RR Lyrae stars and main sequence stars and constrain the metallicity distribution of the Sgr tidal stream. Our main results are summarized and discussed in \\S~\\ref{discussion}. ", "conclusions": "} We have presented the most complete sample of RR Lyrae stars identified so far in the SDSS stripe 82 data set, consisting of 379 RRab and 104 RRc stars. Our visual inspection of single-band and color light curves insures that the sample contamination is essentially negligible. Although the sky area is relatively small compared to other recent surveys, such as \\citet{kel08} and \\citet{mic08}, this RR Lyrae sample has the largest distance limit to date ($\\sim$100 kpc). In addition, a large number of well-sampled light curves in the $ugriz$ photometric system have enabled the construction of a new set of empirical templates. This multi-band empirical template set provides strong constraints for models of stellar pulsations, through the single-band light curve shape, the shape variation with wavelength, and through the distributions of the templates. For example, while $ab$-type RR Lyrae appear to have a continuous distribution of light curve shapes, $c$-type RR Lyrae exhibit a clear bimodal distribution. This bimodality suggests that the latter class may include two distinct subclasses, which would be an unexpected result given that RR Lyrae stars have been carefully studied for over a century (e.g., \\citealt{bai03}). A subclass of RR Lyrae stars with amplitudes and periods similar to the $c$-type, but with asymmetric light curves, has been noted in the MACHO data by \\citet{alc96}. They suggested that these stars are pulsating in the second-harmonic mode. Our accurate templates will greatly assist the identification of RR Lyrae stars in data sets with relatively small numbers of epochs. For example, all three major upcoming wide-area ground-based surveys plan to obtain fewer than 10 epochs (SkyMapper, Dark Energy Survey and Pan-STARRS1 3$\\pi$ survey), in the same photometric system. Without template fitting, the selection based on only low-order light-curve statistics includes a significant fraction ($\\sim$30\\%) of contaminants (dominated by $\\delta$ Scuti stars). The high level of completeness and low contamination of our resulting sample, as well as the more precise distance estimates, enabled a more robust study of halo substructure than was possible in our first study. Our main result remains: the spatial distribution of halo RR Lyrae at galactocentric distances 5--100 kpc is highly inhomogeneous. The distribution of $\\rho/\\rho^{RR}_{model}$ suggests that at least 20\\% of the halo within 30 kpc is found in apparently real substructures. \\citet{bel08} find a lower limit of 40\\%, but they used a different metric and different tracers (turn-off stars). Schlaufman et al.~(in prep) have demonstrated that at least 34\\% of inner-halo (in their paper they explored regions up to 17 kpc from the Galactic center) main sequence turnoff stars are contained in structures they refer to as ECHOS (elements of cold halo substructures), based on an examination of radial velocities in the SEGUE sample. Thus, three separate studies, using very different probes, have come to essentially identical conclusions. A comparison of the observed spatial distribution of RR Lyrae stars and main sequence stars to the \\citet{jur08} model, which was constrained by main sequence stars at distances up to 20 kpc, strongly suggests that the halo stellar number density profile steepens beyond $\\sim$30 kpc. While various indirect evidence for this behavior, based on kinematics of field stars, globular clusters, and other tracers has been published (\\citealt{car07}; and references therein), our samples provide a direct measurement of the stellar halo spatial profile beyond the galactocentric distance limit of $\\sim$30 kpc. A similar steepening of the spatial profile was detected using candidate RR Lyrae stars by \\citet{kel08}. In addition to presenting new evidence for the steepening of halo stellar number density profile beyond $\\sim$30 kpc, we have confirmed the result of \\citet{mic08} that the density profile for Oosterhof II stars is steeper than for Oosterhof I stars within 30 kpc. We have analyzed several methods for estimating the metallicities of RR Lyrae stars. Using spectroscopic data processed with the SDSS SSPP pipeline as a training sample, we have established a weak correlation with the $u-g$ color at minimum light. The rms scatter around the best-fit relation is $\\sim0.3$ dex, but systematic errors could be as large, and the best-fit relation should be used with care. We have established a relationship between the parametrization of our template set for $ab$-type RR Lyrae stars and the results based on the Fourier expansion of light curves. This allowed us to estimate the metallicity from observed light curves with an estimated rms scatter of $\\sim0.3$ dex. This method failed to uncover a systematic metallicity difference between field RR Lyrae stars and those associated with the Sgr stream, which is seen in main sequence stars. Nevertheless, it is likely that this is a promising method for estimating metallicity, which is currently limited by the lack of a large and reliable calibration sample. Our templates can be used to bypass noisy Fourier transformation of sparsely sampled data, as already discussed by \\citet{kin06} and \\citet{kk07}. We introduced a novel method for estimating metallicity that is based on the absolute magnitude vs.~metallicity relations for RR Lyrae stars and main sequence stars (calibrated using globular clusters). This method does not require the $u$-band photometry, and will be useful to estimate metallicity of spatially coherent structures that may be discovered by the Dark Energy and Pan-STARRS surveys. While the existing SDSS data are too shallow to apply this method to the Pisces stream, we used it to obtain a metallicity estimate for the Sgr tidal stream that is consistent with an independent estimate based on the photometric $u$-band method for main sequence stars. Our result, $[Fe/H]=-1.2$, with an uncertainty of $\\sim$0.1 dex, strongly rules out the hypothesis that the trailing arm of the Sgr dSph tidal stream has the same metallicity as halo field stars ($[Fe/H]=-1.5$), as suggested by \\citet{wat09}. \\citet{yan09} detected peaks at $[Fe/H]=-1.3$ and at $[Fe/H]=-1.6$ using SDSS spectroscopic metallicities for blue horizontal-branch (BHB) stars from the trailing arm. However, both the Yanny et al.~and Watkins et al.~results could be affected by systematic errors in the SDSS metallicity scale for BHB stars. Our results suggest that the SDSS metallicity scale for BHB stars could be biased low by about 0.3 dex (relative to the SDSS metallicity scale for main sequence stars, and assuming that RR Lyrae and main sequence stars from the Sgr tidal stream in stripe 82 area have the same metallicity distributions). Simulations by \\citet{bj05} and \\citet{joh08} predict that there should be a difference in the chemical composition between stars in the inner halo that was built from accretion of massive satellites about 9 Gyr ago, and outer halo dominated by stars coming from dSph satellites that were accreted in the last 5 Gyr (see Fig.~11 in \\citealt{bj05}). These accreted dSph satellites were presumably more metal-poor than the massive satellites accreted in earlier epochs (see Fig.~3 in \\citealt{rob05}). Other recent simulation studies support these conclusions. For example, \\citet{dh08} find evidence in their simulation for a strong concentration of (relatively) higher metallicity stars at distances close to the Galactic center, and the presence of (relatively) lower metallicity stars at distances beyond 20 kpc from the center. \\citet{zol09} find from their simulations that their inner halos include stars from both in-situ formed stars and accreted populations, while their outer halos appear to originate through pure accretion and disruption of satellites. Salvadori et al.~(in prep) have considered the distribution of ages and metallicities of metal-poor stars in a Milky-Way like halo, as a function of galactocentric radius, based on a hybrid N-body and semi-analytic simulation. They find an inner-halo population that is well-described by a power-law index $n=-3.2$ (for stars with $-2<[Fe/H]<-1$), and an outer-halo consistent with a much shallower profile, $n=-2.2$. The relative contributions of stars with $[Fe/H]\\le-2$ in their simulation increases from about 16\\% for stars within 7 kpc $< R <$ 20 kpc, to $>40\\%$ for stars with $R >$ 20 kpc. Our observational results provide some support for these simulation-based predictions. The faintest main sequence stars ($r\\sim21.5$), not including those apparently associated with the Sgr stream, exhibit median metallicities at least 0.3 dex lower ($[Fe/H]=-1.8$, see the left panel in Figure~\\ref{fig:bsFig3}) than halo stars within 10-15 kpc from the Galactic center \\citep{ive08}. Other studies provide additional support. For example, \\citet{car07} and Carollo et al.~(in prep) have indicated that field stars likely to be associated with the outer halo exhibit metallicities that are substantially lower ($[Fe/H]=-2.2$) than those of the inner halo (they place the inner/outer halo boundary at $\\sim$10 kpc). The two metallicity measurements are fully consistent because photometric $u$-band metallicity estimator is biased high for spectroscopic values with $[Fe/H]< -2$ (Bond et al., in prep). Carollo et al.~(in prep) also found that the inferred density profiles of the inner- and outer-halo populations differ as well; the inner halo being consistent with a power-law profile with index $n \\sim -3.4$, and the outer halo having index $n = -2.1$ (the data analysed by Carollo et al. are not suitable for determination of the relative normalizations of the inner- and outer-halo populations, due to the selection effects involved). This difference in the density profiles is very similar to the difference in density profiles for Oosterhof I and Oosterhof II stars. However, we emphasize that the steeper profile for the inner halo from Carollo et al.~corresponds to more metal-rich stars, while we found weak evidence that Oosterhof II stars tend to be more metal poor than Oosterhof I stars. Simulations indicate that high surface brightness substructures in the halo originate from single satellites, typically massive dSph which tend to be accreted over the last few Gyr \\citep{bj05}, and these massive galaxies are expected to be more more metal-rich than halo field stars \\citep{fon08}. The results from \\citet{ive08} and the results presented here seem to support this prediction. The inner halo has a median metallicity of $[Fe/H]=-1.5$, while at least two strong overdensities have higher metallicities -- the Monoceros stream has $[Fe/H]=-1.0$, and for the trailing part of the Sgr tidal stream we find $[Fe/H]=-1.2$. We emphasize that these three measurements are obtained using the same method/calibration and the same data set, and thus the measurements of relative differences are expected to be robust. Simulations by \\citet{joh08} predict that the metallicity of low surface brightness features, such as the Virgo Overdensity \\citep{jur08}, is expected to be lower than the median halo metallicity. Although the initial estimate by \\citet{jur08} (see their Figure 39) claimed a metallicity for this structure of $[Fe/H]\\sim-1.5$, a recent study of the photometric metallicity of stars in this structure by \\citet{an09} have suggested that the mean metallicity is even lower: $[Fe/H]= -2.0$. Both of these studies are limited by potentially large systematics associated with the photometric metallicity technique, so more detailed analysis of spectroscopically determined metallicities for stars in this structure would be of great value. The sparse information that is available, based on spectroscopic determinations, supports a mean metallicity between $[Fe/H]= -1.8$ and -2.2 \\citep{duf06,viv08,pri09,cas09}. \\citet{cas09} report the measurement of a precise absolute proper motion for the RR Lyrae star RR 167, which appears highly likely to be associated with the Virgo Overdensity. This proper motion, in combination with their distance estimate (17 kpc from the Galactic center) and radial velocity measurement, indicate that the Virgo Overdensity structure may well be on a highly destructive orbit, with pericenter $\\sim 11$ kpc, and apocenter $\\sim 90$ kpc. Thus, the interpretation of the Virgo Overdensity as a dwarf galaxy, perhaps similar to those that participated in formation of the outer-halo population, is strengthened. Our result that (inner-) halo stellar number density profile steepens beyond $\\sim$30 kpc is limited by the relatively small distance limit for main sequence stars (35 kpc), the sparseness of the RR Lyrae sample ($\\sim$500 objects), and the small survey area ($\\sim$300 deg$^2$). Ideally, the halo stellar number density profile should be studied using numerous main sequence stars detected over a large fraction of sky. To do so to a distance limit of 100 kpc, imaging in at least $g$ and $r$ bands (or their equivalent) to a depth several magnitudes fainter than the co-added SDSS stripe 82 data is required ($r>25$). Pan-STARRS, the Dark Energy Survey and LSST are planning to obtain such data over large areas of sky. The LSST, with its deep $u$-band data, will also extend metallicity mapping of field main sequence stars over half of the sky in the south; see \\citet{ive08} for details. For substructures to be potentially discovered in the north by Pan-STARRS, the method presented here can be used to estimate the metallicity of spatially coherent structures even without the $u$-band data." }, "0910/0910.1728_arXiv.txt": { "abstract": "The correlation between the large-scale distribution of galaxies and their spectroscopic properties at $z=1.5$ is investigated using the Horizon MareNostrum cosmological run. We have extracted a large sample of $10^5$ galaxies from this large hydrodynamical simulation featuring standard galaxy formation physics. Spectral synthesis is applied to these single stellar populations to generate spectra and colours for all galaxies. We use the skeleton as a tracer of the cosmic web and study how our galaxy catalogue depends on the distance to the skeleton. We show that galaxies closer to the skeleton tend to be redder, but that the effect is mostly due to the proximity of large haloes at the nodes of the skeleton, rather than the filaments themselves. This effects translate into a bimodality in the colour distribution of our sample. The origin of this bimodality is investigated and seems to follow from the ram pressure stripping of satellite galaxies within the more massive clusters of the simulation. The virtual catalogues (spectroscopical properties of the MareNostrum galaxies at various redshifts) are available online at {\\tt http://www.iap.fr/users/pichon/} {\\tt MareNostrum/catalogues}. ", "introduction": "\\label{sec:intro} \\begin{figure} \\includegraphics[width= 1.1\\columnwidth]{fig/blah} \\caption{A 3D view of the galaxies and the skeleton of the dark matter of MareNostrum at $z=1.6$. The box is 50 $h^{-1}$Mpc aside.} \\label{fig:skel} \\end{figure} During the past decade, the $\\Lambda {\\rm {CDM}}$ cosmological model of the Universe has been established as the framework of choice in which to interpret how and when observed galaxies acquire their properties. Arguably the most important feature of this framework is to provide us with an explanation as to why many of these properties (physical sizes, luminosities) strongly correlate with galaxy mass while others (star formation rates, morphological type) do not seem to. Unsurprizingly, the all-time favored culprit is the interplay between galaxies and the intergalactic medium (IGM) at large. In other words, the large scale environment of galaxies is claimed to play an important role in shaping some of their properties, while the rest of them are thought to depend solely on small scale (internal) processes. However, having said that, one still has to determine for which of these properties ``Nurture\" dominates over ``Nature\" and therein lies the whole difficulty of the issue. Indeed, since the early 70s, there has been a plethora of studies devoted to measuring the impact of environment on galaxy properties. \\cite{DavisGeller1976} first pointed out that early-type galaxies are more strongly clustered than the late types, while \\cite{Dressler1980} and \\cite{PostmanGeller1984} demonstrated the existence of a morphology-density relation (MDR). Following in their footsteps, \\cite{Balogh1998}, \\cite{Hashimoto} and more recently \\cite{Christlein} and \\cite{Poggianti} systematically showed that galaxies living in denser environments tend to be redder and have lower star formation rates (SFRs) than their more isolated counterparts. One can think of several physical processes associated with different types of environment that could play a role in causing such alterations. More specifically, they include, by ascending order of environment density: (i) major (wet) mergers, which can turn spiral galaxies into ellipticals (e.g. \\cite{toomre}), and drive a massive starburst wind which quenches future star formation by ejecting the interstellar medium (ISM) out of galaxies (\\cite{Mihos,maclow}); (ii) active galactic nuclei (AGN) or shock-driven winds (\\cite{norm,sprin}; (iii) galaxy ``harassment\" (rapid encounters or fly-bys which dominate over mergers in rich clusters) causing discs to heat and possibly triggerring the build-up of a bulge via the formation of a bar (\\emph{e.g.} \\cite{moore}). For the latter category, the diffuse gas associated with the galaxies' host dark matter (sub)halo which constitutes the main fuel supply for future star formation can also be stripped, thus suppressing later star formation by ``starvation\" or ``strangulation\" \\citep{larson,Bekki}. Moreover, part of their ISM can also be pulled out of these galaxies, either by tidal forces arising from the gravitational potential of the cluster or by ram pressure stripping by the intracluster medium (ICM) \\cite{Gunn,Abadi,Chung}. \\\\ Although these latter environment dependent processes seem potent enough, recent work carried out by \\cite{Tanaka} and \\cite{vdB} indicate that they might not be the main mechanisms for quenching star formation activity. This claim is corroborated by the higher redshift results ($ z \\!\\sim\\! 1$) obtained with the DEEP2 (e.g. \\cite{Cooper2006,Cooper2007,Cooper2008,Gerke2007,Coil2008}), (z--)COSMOS \\citep{Scoville2007,Cassata2007,Tasca2009} and VVDS \\citep{Scodeggio} surveys where clusters are more scarce, along with the fact that morphological and spectrophotometric properties of local galaxies are also found to be correlated with their internal properties, such as luminosity, mass or internal velocity (e.g. \\cite{Kauffmann2003}). Here as well, one can invoke various physical processes to explain such dependences on internal properties. Supernova feedback can heat and stir the ISM, possibly ejecting large amounts of gas out of galaxies and it is expected to scale with galaxy mass \\citep{Larson1974,Dekel1986}. There also exists a growing host of observational evidence that AGN play a key part in quenching star formation \\citep{Schawinski2006,Schawinski2007,Salim2007} and this AGN feedback should likely impact more massive galaxies since they host larger mass black holes \\citep{Magorrian,Silk1998}. In light of these investigations, it becomes apparent that disentangling nature and nurture is a more complicated process than one would naively have thought to begin with. Clearly, the fundamental requirement to tackle this issue is to properly characterize the anisotropic environment of galaxies, both in observational samples and theoretical models, spanning as broad a range of environments as possible, from isolated field galaxies to groups and rich clusters. The vast majority of the studies in the literature accomplish this task either by counting the number of neighbours that a galaxy has within a fixed aperture on the sky or by measuring the distance to the $n^{\\rm th}$ nearest galaxy, where $n$ is an integer in the range 3\u201310. Although these indicators are straightforward to obtain, their physical interpretation, let alone their comparison to theoretical models are far from being straightforward (\\emph{cf.} \\cite{Kauffmann2004,Weinmann2006}). Meanwhile, looking at the distribution of observed galaxies in modern cosmological surveys, such as the 2dF \\citep{2dF} and the SDSS \\citep{SDSS}, the most striking feature is that they look organised along linear structures linking clusters together (see Figure \\ref{fig:skel}). This filamentary network, dubbed as the ``Cosmic Web\" \\citep{bond}, has a dynamical origin and reflects the anisotropic accretion taking place in clusters \\citep{skel1}. It therefore seems natural to describe the environment of galaxies in terms of their location with respect to these filaments in order to investigate the influence of the Cosmic Web on the properties of the galaxies it encompasses. In this paper, we carry out such a study at intermediate redshift ($z\\sim1.5$), mainly from a theoretical perspective, using a recent diagnostic tool to characterize the 3D environment called the skeleton \\citep{sousb08}, which we combine with the largest hydrodynamical cosmological simulation performed to date \\citep{ocvirk2008,prunet2008, dekelnature}. Run on the MareNostrum computer at the Barcelona Supercomputer Center using the RAMSES code \\citep{teyssier02}, this simulation is one of the flagship simulations realized by the Horizon collaboration ({\\tt http://www.projet-horizon.fr}). It includes a detailed treatment of metal--dependent gas cooling, UV heating, star formation, supernovae feedback and metal enrichment. Specifically, we will address the question: are the physical conditions \\emph{within the filaments} dramatic enough to strongly influence the properties of the galaxies it encompasses? The outline of this paper is as follows: first we describe in Section \\ref{sec:method} our methodology, in terms of numerical techniques, estimators and statistical measurements. The dependence of the spectroscopic properties on the filamentary environment is then discussed in Section \\ref{sec:fil}, while Section \\ref{sec:bimodality} investigates the observed bimodality and discusses comparison to observations, and Section \\ref{sec:conclusion} wraps up. Some checks are performed in Appendix A, Appendix B describes the publicly available catalogues and Appendix C sums up the subgrid physics used. ", "conclusions": "\\label{sec:conclusion} The Cosmic Web is a key feature of the organisation of galaxies on large scales. This paper investigated the influence of this filamentary environment on the spectroscopic properties of galaxies, using the MareNostrum simulation which was postprocessed using stellar population synthesis. The cosmic web was traced with the skeleton algorithm, and has proven here to be a very effective mean of probing the anisotropy of the large-scale structures. We found gradients of spectroscopic properties of galaxies with the distance to filaments, but demonstrated that they can be explained by the fact that the distance to filaments is biased by the distance to the nodes of the network (group or clusters of galaxies). Two procedures were introduced to remove the influence of these nodes: (i) volume-averaging this distance decreases the influence of the compact, dense regions and focus on wider structures, such as filaments; (ii) pair comparison which seeks the filamentary counterpart of each galaxy. Both methods show that the influence of the filaments alone is negligible compared to influence of clusters. A bimodality in colour was also found to occur below redshift $z\\approx 2$ and its origin was investigated. It is due to a population of red, small galaxies ($\\sim 10^8 M_\\odot$) accreting on the nodes of the Cosmic Web, while more massive objects, $M> 10^8 M_\\odot$, are mostly unaffected. These galaxies have their star formation quenched while they approach the clusters. It remains to be confirmed that this stripping process is not amplified by a lack-of-resolution effect, since (i) it involves amongst the smallest (virtual) galaxies in the simulation, and (ii) it seem to create a tension with observations at redshift zero \\citep{kimm}. These findings suggest that the large-scale filaments are only dynamical features of the density field, reflecting the flow of galaxies accreting on clusters; the conditions in the filaments are not dramatic enough to influence strongly the properties of the galaxies it encompases, unlike the intra (proto) cluster medium, which seems able to strip down the ISM of the incoming low mass galaxies. Appendix \\ref{sec:test_cut} shows that this statement remains valid when the study is limited to the most important filaments. Finally, we did find a weak metallicity gradient away from the filaments which could reflect the large-scale inhomogeneity in the distribution of metals (the so called WHIM). One could have imagined that even if the large-scale filamentary network were purely a tracer of the large-scale dynamics, galaxies within the large-scale filaments should be redder and older, since they would have joined the cosmic super highway earlier on average when compared to field galaxies. This effect is not seen at those redshifts, as (i) older galaxies continue to accrete new cold gas on small scales and form stars, hence remain blue, (ii) some field galaxies continuously join the large filamentary network, and (iii) on large-scales, a significant fraction of the matter is accreted more or less radially onto the nodes, while filaments are collecting left-overs from this accretion more indirectly \\citep{skel1}. Recently \\citep{keres05,ocvirk2008,dekelnature}, it was emphasized in steps that anisotropic metal-rich cold stream accretion regulates the inflow of cold gas towards the inner regions of the most massive galaxies of the high redshift universe ($z>2$). It was then conjectured that this process could explain the observed bimodality of spectroscopic galactic properties at lower redshift. In this paper, we have shown that the geometric distribution of the colour of galaxies is not sensitive to the detailed large-scale filamentary network, but only to its nodes. This apparent paradox may be lifted when noting that the self-regulating anisotropic filamentary accretion occurs on much smaller scales, and was quantified for the central galaxies of the simulation which {\\sl are} sitting at the nodes of the network; in contrast, when considering the full galactic population, a typically low mass galaxy is not transformed by its encounter with the different physical condition of the weakly overdense intergalactic medium within the cosmic web, unless it falls into the intra cluster medium of a large node.\\\\ In other words, massive galaxies feel the small scale filaments (cold streams) feeding them at nodes; low mass galaxies are not spectroscopically changed while entering the large-scale filaments. The mesoscopic (below a Mpc scale) filamentary feature of the cosmic network may geometrically solve the self-regulating process of galactic accretion, but we have demonstrated here that its large-scale counterpart does not seem to directly affect the colours of galaxies. In \\cite{sousb08}, the effect of redshift distortion is partially addressed, and the corresponding algorithm is now being extended to discrete surveys via a Delaunay tessellation (which are therefore not sensitive to edge effects). This, as argued in Section~\\ref{sec:discussion} is a critical step towards performing a similar 3D analysis on real data, which as we have shown is essential to quantify these gradients, as in projection, the information is lost (see Figure~\\ref{fig:2D}). In particular, it would be of great interest to carry out these measurements on a DEEP-2/VVDS/z-COSMOS-like survey as well as to bring the simulation down to lower redshift and reach a time in cosmic history when upcoming large observational surveys (\\emph{e.g.} LSST \\cite{LSST}, BOSS \\cite{BOSS}) overlap statistically with the predictions of the simulation. It would also be worth investigating how sensitive some of our findings are w.r.t. the detailed chosen subgrid physics by probing alternative recipes and running higher spatial resolution simulations. The catalogues produced for this investigation (spectroscopical properties of the MareNostrum galaxies and its dark matter skeletons) are available online as discussed briefly in Appendix~\\ref{sec:catalogs}. \\subsection*{Acknowledgements} { \\sl We thank F.~Brault for her help in a preliminary investigation, S.~Colombi, D.~Pogosyan, Y.~Dubois and D.~Aubert for fruitful comments during the course of this work. This investigation was carried within the framework of the Horizon project, \\texttt{www.projet-horizon.fr}. The simulation was run on the MareNostrum machine at the Barcelona Supercomputing Centre and we would like to warmly thank the staff for their support and hospitality. We also thank D.~Munro for freely distributing his Yorick programming language and opengl interface (available at {\\em\\tt http://yorick.sourceforge.net/}). CP thanks the Leverhulme Trust for the visiting professorship F09846D. }" }, "0910/0910.4427_arXiv.txt": { "abstract": "We report the results of our observations of H{\\sc i} absorption towards the central region of the rejuvenated radio galaxy 4C29.30 (J0840+2949) with the Giant Metrewave Radio Telescope (GMRT). The radio source has diffuse, extended emission with an angular size of $\\sim$520 arcsec (639 kpc) within which a compact edge-brightened double-lobed source with a size of 29 arcsec (36 kpc) is embedded. The absorption profile which is seen towards the central component of the inner double is well resolved and consists of six components; all but one of which appears to be red-shifted relative to the optical systemic velocity. The neutral hydrogen column density is estimated to be $N$(H{\\sc i})=4.7$\\times$10$^{21}$($T_s$/100)($f_c$/1.0) cm$^{-2}$, where $T_s$ and $f_c$ are the spin temperature and covering factor of the background source respectively. This detection reinforces a strong correlation between the occurrence of H{\\sc i} absorption and rejuvenation of radio activity suggested earlier, with the possibility that the red-shifted gas is fuelling the recent activity. ", "introduction": "One of the interesting and important questions in our understanding of active galactic nuclei (AGN) is whether such activity is usually episodic, and if so, the range of time scales of AGN activity. Besides helping to constrain models of episodic acitvity, this also has wider implications in our understanding of AGN feedback in structure formation and the evolution of galaxies (e.g. Sijacki et al. 2007, and references therein; Nesvadba \\& Lehnert 2008). In the presently widely accepted paradigm, AGN activity is believed to be intimately related to the `feeding' of a supermassive black hole whose mass ranges from $\\sim$10$^6$ to 10$^{10}$ M$_\\odot$ (e.g. Marconi et al. 2004). Periodic `feeding' of the supermassive black hole may lead to different cycles of activity. \\begin{figure*} \\vbox{ \\psfig{file=VLA1400collage.ps,width=6.35in,angle=00} } \\caption[]{A collage showing the large- and small-scale structure of the rejuvenated radio galaxy 4C29.30 (J0840+2949) at 1400 MHz reproduced from Jamrozy et al. (2007). The left panel shows the D-array contour map of the entire source overlayed on the optical field from the Digital Sky Survey (DSS). The contour levels are spaced by factors of $\\sqrt{2}$ and the first contour is 0.3~mJy~beam$^{-1}$. The right panel shows the contour map of the central part of the source, containing the inner double, from the FIRST (Faint Images of the Radio Sky at Twenty-cm, Becker, White \\& Helfand 1995) survey. The contour levels are spaced by factors of $\\sqrt{2}$, and the first contour is 0.45~mJy~beam$^{-1}$. The size of the beam is indicated by an ellipse in the bottom right corner of each image. } \\end{figure*} For radio-loud AGN, an interesting way of probing their history and hence episodic jet activity is via the structural and spectral information of the lobes of extended radio emission. (e.g. Burns, Schwendeman \\& White 1983; Burns, Feigelson \\& Schreier 1983; van Breugel \\& Fomalont 1984; Leahy, Pooley \\& Riley 1986; Baum et al. 1990; Clarke, Burns \\& Norman 1992; Junkes et al. 1993; Roettiger et al. 1994; Schoenmakers et al. 2000; Gizani \\& Leahy 2003; Konar et al. 2006). A very striking example of episodic jet activity is when a new pair of radio lobes is seen closer to the nucleus before the `old' and more distant radio lobes have faded (e.g. Subrahmanyan, Saripalli \\& Hunstead 1996; Lara et al. 1999). Such sources have been christened as `double-double' radio galaxies (DDRGs) by Schoenmakers et al. (2000). Saikia, Konar \\& Kulkarni (2006) reported the discovery of a new DDRG J0041+3224 and compiled a sample of approximately a dozen such objects from the literature, including 3C236 (Schilizzi et al. 2001) and J1247+6723 (Marecki et al. 2003; Bondi et al. 2004). The inner doubles in these two sources are compact with sizes of 1.7 kpc and 14 pc respectively, and have been classified as a compact steep spectrum (CSS) and a Gigahertz peaked spectrum (GPS) source respectively. The median linear size of the inner doubles for this sample of DDRGs is $\\sim$150 kpc, while the overall median total linear size is approximately a Mpc. In addition to the classic DDRGs, evidence of episodic activity may also be seen as diffuse, steep-spectrum emission from an earlier cycle of activity, in which a young double-lobed radio source may be embedded. An archetypal example of such a source, namely 4C29.30, was studied in detail by Jamrozy et al. (2007). If the nuclear or jet activity is rejuvenated by a fresh supply of gas one might be able to find evidence of this gas via H{\\sc i} absorption towards the radio components in the central regions of the host galaxy. Saikia, Gupta \\& Konar (2007) reported the detection of H{\\sc i} absorption towards the inner double of the DDRG, J1247+6723. From the available information in the literature, they also suggested that there could be a strong relationship between the detection of H{\\sc i} gas and rejuvenation of radio activity. We present the results of H{\\sc i} absorption towards the rejuvenated radio galaxy 4C29.30 with the Giant Metrewave Radio Telescope (GMRT). The total flux density of the inner double at 1287 MHz was estimated by Jamrozy et al. (2007) to be 390 mJy when observed with an angular resolution of $\\sim$2.6 arcsec, suggesting it to be a suitable source for these observations. We describe some of the basic properties of 4C29.30 (J0840+2949) in Section 2, the observations and analyses in Section 3 and present the results and discussions in Section 4. The conclusions are summarised in Section 5. ", "conclusions": "Our GMRT image of the source (Fig. 2) with an angular resolution of 3.60$\\times$2.35 arcsec$^2$ along a position angle of 45.7$^\\circ$ and an rms noise of 0.4 mJy beam$^{-1}$ detects the inner double with a peak brightness of 78.9 mJy beam$^{-1}$ for the central component and 74.6 and 71.5 mJy beam$^{-1}$ for the northern and southern hotspots of the inner double. The peak brightness of the knot in the jet towards the south is 15.8 mJy beam$^{-1}$. The H{\\sc i} absorption spectrum towards the central component is presented in Fig. 3. H{\\sc i} absorption has been detected clearly towards the core of this rejuvenated radio galaxy. The rms noise in the spectrum is $\\sim$1 mJy beam$^{-1}$ channel$^{-1}$. The absorption profile consists of a number of components all but one of which appear red-shifted relative to the optical systemic velocity. The H{\\sc i} column density, $N$(H{\\sc i}), for the different spectral components has been estimated using the relation \\begin{equation} $N$({\\rm H{\\sc i}})=1.93\\times10^{18}\\frac{{T}_{s}~\\tau_p~\\Delta v}{f_c}~ {\\rm cm^{-2}}, \\label{eq1} \\end{equation} where $T_s$, $\\tau_p$, $\\Delta$$v$ and $f_c$ are the spin temperature, peak optical depth, the full width at half maximum (FWHM) of the Gaussian line profile, and the fraction of background emission covered by the absorber respectively. The estimates have been made assuming $T_s$=100 K and $f_c$=1.0. The value of $T_s$ could be significantly greater than 100 K. For example, for the warm neutral medium seen in the Galaxy $T_s$ ranges from 5000$-$8000 K (Kulkarni \\& Heiles 1988). Such high spin temperatures are also expected to arise in the proximity of an active nucleus (Bahcall \\& Ekers 1969; Holt et al. 2006). The best fit to the spectrum with six Gaussian components is shown in Fig. 2 and the fit parameters are summarised in Table 1. The sixth component requires confirmation from more sensitive observations. The Table lists the optical heliocentric velocity, v$_{\\rm{hel}}$, FWHM of the Gaussian profile, the fractional absorption and the H{\\sc i} column density. The numbers within brackets are the errors on the quoted values. The H{\\sc i} column density integrated over the entire absorption profile is 4.7$\\times$10$^{21}$ cm$^{-2}$. The H{\\sc i} absorption spectra towards the hotspots and the `southern knot' are shown in Fig. 4. No significant absorption has been detected towards the northern and sourthern hotspots of the inner double or the `southern knot' indicating a 3-$\\sigma$ upper limits to H{\\sc i} of $N$(H{\\sc i})=9.5$\\times$10$^{20}$ cm$^{-2}$, $N$(H{\\sc i})=8.3$\\times$10$^{20}$ cm$^{-2}$ and $N$(H{\\sc i})=5.5$\\times$10$^{21}$ cm$^{-2}$ respectively, assuming $T_s$=100 K, $f_c$=1.0 and $\\Delta$$v$=100 km s$^{-1}$. This indicates that the size of the absorber is less than $\\sim$35 kpc. However, the spectra of the `southern knot' and the northern hotspot show very marginal signs of absorption at $\\sim$1.948$\\times$10$^4$ km s$^{-1}$ which needs to be confirmed from more sensitive observations. \\subsection{H{\\sc i} gas and rejuvenation of AGN activity} Saikia et al. (2007) explored any possible relationship between rejuvenation of radio or jet activity and the occurrence of H{\\sc i}. Unfortunately the number of sources is still small because most of the rejuvenated radio sources have weak radio emission in the central or nuclear region. In the list of DDRGs compiled by Saikia et al. (2006) the two exceptions are 3C236 and J1247+6723, with the flux density within a few kpc of the nuclear region being $\\gapp$100 mJy. The DDRG 3C236 which is a giant radio galaxy with a projected linear size of $\\sim$4250 kpc shows evidence of star formation and H{\\sc i} absorption against a lobe of the inner radio source (Conway \\& Schilizzi 2000; Schilizzi et al. 2001; O'Dea et al. 2001). Observations with milliarcsec resolution are required to determine the location of the absorbing clouds in the case of J1247+6723 reported by Saikia et al. (2007). An interesting case is 3C293, which could be classified as a misaligned DDRG, and exhibits absorption features both blue- and red-shifted relative to the the systemic velocity (Beswick, Pedlar \\& Holloway 2002; Beswick et al. 2004), and fast outflowing gas blue-shifted by upto $\\sim$1000 km s$^{-1}$ (Morganti et al. 2003; Emonts et al. 2005). The case for 3C258 (J1124+1919) as a rejuvenated galaxy is less clear. Although H{\\sc i} absorbing gas was reported by Gupta et al. (2006) towards the compact central source (e.g. Sanghera et al. 1995), we have not been able to confirm from GMRT observations (Saikia et al., in preparation) the weak extended emission reported by Strom et al. (1990). Another celebrated case is Centaurus A, where the inner double is due to more recent activity while the extended diffuse emission is due to earlier cycles of nuclear activity (Burns et al. 1983; Junkes et al. 1993; Ilana Feain, private communication), and H{\\sc i} absorption is seen towards the central region (e.g. Sarma, Troland \\& Rupen 2002; Morganti et al. 2008). While the number of sources is still small, the detection of absorbing H{\\sc i} gas in the rejuvenated galaxies appears to be more frequent than for CSS and GPS objects (Vermeulen et al. 2003; Gupta et al. 2006), and the reported observations of 4C29.30 reinforces this trend. Although the sample size needs to be clearly increased, this trend is unlikely to be due to different source strengths. Considering the GPS objects listed by Gupta et al. (2006), which has the highest H{\\sc i} detection rate of $\\sim$45 per cent, the median total flux density of the sources at $\\sim$1400 MHz is $\\sim$2 Jy. Amongst the rejuvenated galaxies discussed here, the flux densities at $\\sim$1400 MHz of the entire central source of 3C236, 3C293 and Cen A are comparable to those of the GPS objects, while those of 4C29.30 and J1247+6723 are weaker than the GPS objects listed by Gupta et al. (2006). There should not be any bias because of source strength. Considering the H{\\sc i} column densities (Gupta et al. 2006), the median value for GPS sources is $\\sim$3$\\times$10$^{20}$ cm$^{-2}$. The rejuvenated radio galaxies discussed here have column densities in the range of $\\sim$8$-$50$\\times$10$^{20}$ cm$^{-2}$. \\begin{table} \\caption{Multiple Gaussian fit to the H{\\sc i} absorption spectrum towards the central component (Fig. 3).} \\begin{center} \\begin{tabular}{|c|l|l|c|c|} \\hline Id. & v$_{\\rm{hel}}$ & FWHM & Frac. abs. & $N$(H{\\sc i}) \\\\ no. & & & & 10$^{20}$ \\\\ % & km s$^{-1}$ & km s$^{-1}$ & & cm$^{-2}$ \\\\ \\hline 1 & 19383(2) & 37(5) & 0.072(0.008) & 5.4(1.4) \\\\ 2 & 19457(2) & 33(3) & 0.222(0.009) & 16.1(2.0) \\\\ 3 & 19490(1) & 25(3) & 0.243(0.013) & 13.4(2.1) \\\\ 4 & 19519(1) & 18(2) & 0.238(0.018) & 9.3(1.6) \\\\ 5 & 19549(4) & 36(11) & 0.062(0.009) & 4.4(2.0) \\\\ 6 & 19599(1) & 14(4) & 0.065(0.013) & 1.8(1.0) \\\\ \\hline \\end{tabular} \\end{center} \\label{gauss} \\end{table}" }, "0910/0910.1034_arXiv.txt": { "abstract": "{ Observations of dense molecular gas lie at the basis of our understanding of the density and temperature structure of protostellar envelopes and molecular outflows. The Atacama Pathfinder EXperiment (APEX) opens up the study of southern (Dec $<$ -35$^{\\circ}$) protostars. } {We aim to characterize the properties of the protostellar envelope, molecular outflow and surrounding cloud, through observations of high excitation molecular lines within a sample of 16 southern sources presumed to be embedded YSOs, including the most luminous Class I objects in Corona Australis and Chamaeleon.} {Observations of submillimeter lines of CO, HCO$^+$ and their isotopologues, both single spectra and small maps (up to $80''\\times80''$), were taken with the FLASH and APEX-2a instruments mounted on APEX to trace the gas around the sources. The HARP-B instrument on the JCMT was used to map IRAS 15398-3359 in these lines. HCO$^+$ mapping probes the presence of dense centrally condensed gas, a characteristic of protostellar envelopes. The rare isotopologues C$^{18}$O and H$^{13}$CO$^+$ are also included to determine the optical depth, column density, and source velocity. The combination of multiple CO transitions, such as 3--2, 4--3 and 7--6, allows to constrain outflow properties, in particular the temperature. Archival submillimeter continuum data are used to determine envelope masses.} {Eleven of the sixteen sources have associated warm and/or dense ($\\geq 10^6$ cm$^{-3}$) quiescent gas characteristic of protostellar envelopes, or an associated outflow. Using the strength and degree of concentration of the HCO$^+$ 4--3 and CO 4--3 lines as a diagnostic, five sources classified as Class I based on their spectral energy distributions are found not to be embedded YSOs. The C$^{18}$O 3--2 lines show that for none of the sources, foreground cloud layers are present. Strong molecular outflows are found around six sources, with outflow forces an order of magnitude higher than for previously studied Class I sources of similar luminosity. } {This study provides a starting point for future ALMA and Herschel surveys by identifying truly embedded southern YSOs and determining their larger scale envelope and outflow characteristics.} ", "introduction": "During the early stages of low-mass star formation, Young Stellar Objects (YSOs) are embedded in cold dark envelopes of gas and dust, which absorb the radiation from the central star \\citep{Lada87,Andre93}. This extinction is strong enough that low-mass embedded YSOs, or protostars, only emit weakly at infrared wavelengths \\citep[e.g.,][]{Jorgensen05,Gutermuth08}. Only at later evolutionary phases, in which the envelope has been accreted and/or dispersed, does emission in the optical and infrared (IR) dominate the Spectral Energy Distribution (SED) \\citep[e.g.][]{Hartmann05}. Protostars emit the bulk of their radiation at far-IR and sub-millimeter wavelengths, both as continuum radiation, produced by the cold ($T<30$ K) dust, and through molecular line emission from the gas-phase species present throughout the protostellar envelope. Although the bulk of the mass is accreted during the earliest embedded phases, more evolved protostellar envelopes still contain a reservoir of gas and dust that can accrete onto the central star and disk system and thus provide the material for disk and planet formation. At the same time, jets and winds from the young star interact with the envelope and drive molecular outflows which clear the surroundings. Characterizing and quantifying all of these different physical components in the protostellar stage is still a major observational challenge. The protostellar envelopes and molecular outflows can be directly observed either through thermal emission of dust at (sub)millimeter wavelengths \\citep[e.g.][]{Shirley00,Johnstone01,Nutter05} or the line emission of molecules. Low frequency molecular emission traces the cold gas in the protostellar envelopes \\citep[e.g.,][]{Hogerheijde98,Jorgensen02,Maret04} or molecular outflows \\citep[e.g.,][]{Snell90,Cabrit92,Bachiller99}. Single-dish observations of dust using current generation bolometer arrays are able to map large areas and image the surroundings of protostars \\citep[e.g.,][]{Motte98,Shirley00,Stanke06,Nutter08}. Through radiative transfer models \\citep[e.g.,][]{Shirley02,Jorgensen02,Young03}, including information from shorter wavelengths \\citep[e.g.,][]{Hatchell07}, the temperature and density structure of the protostellar envelope can be constrained, but the continuum data cannot determine the velocity structure of the infalling envelope, characterize the outflows and their interaction with the surroundings, or disentangle envelope and (foreground) cloud material. Analysis of gas observations in the form of spectra of multiple transitions of the same molecule and its isotopologues provide additional strong constraints on the physical characteristics of the protostellar envelope \\citep{Mangum93,Blake94,vanDishoeck95,Schoeier02,Maret04,Jorgensen05,Evans05,vanderTak07} and outflowing gas \\citep[e.g.,][]{Cabrit92,Bontemps96,Hogerheijde98,Hatchell99,Parise06,Hatchell07a}. Although many different molecules have been observed in protostellar envelopes, only a limited number of species are well suited to trace the physical characteristics of all of the components of an embedded YSO and its surroundings. For example, the use of CH$_3$OH and H$_2$CO is complicated by their changing abundances through the envelope, although some information can be obtained with careful analysis and a sufficient number of observed lines \\citep[e.g.][]{Mangum93, Blake94, vanDishoeck95, vanderTak00,Leurini04}. The weakness of high-excitation CH$_3$OH and H$_2$CO lines in all but a handful of the most luminous Class 0 protostars \\citep[e.g.][]{Jorgensen05b} coupled with a lack of some collisional rate coefficients make these species less suitable tracers for the bulk of the low-mass embedded YSOs. In practice, the column density of the surrounding cloud (and that of any unrelated foreground clouds) is best probed by low excitation optically thin transitions of molecules with a low dipole moment, such as the isotopologues of CO, e.g., C$^{18}$O 2--1 or 3--2 \\citep{vanKempen09}. For molecular outflows the line wings of various $^{12}$CO transitions have been efficiently used as tracers of their properties \\citep{Cabrit92,Bontemps96,Hogerheijde98,Bachiller99,Hatchell99,Hatchell07a}. The denser regions of the protostellar envelope need to be probed with emission lines with a high critical density, such as the higher excitation transitions from HCO$^+$ with critical densities $>10^6$ cm$^{-3}$. The warm gas ($T>$50 K), which can be present in both the molecular outflow and the inner region of the protostellar envelope, can only be traced by spectrally resolved high-$J$ CO transitions, which have $E_{\\rm{up}}>50 $ K for $J_{\\rm{up}}\\geq 4$ \\citep{Hogerheijde98}. Embedded YSOs are generally identified by their 2-24 $\\mu$m IR slope with positive values characteristic of Class I objects \\citep{Lada87}. Over the last several years, it has been found that some of these Class I objects turn out to be edge-on disks or obscured sources \\citep[e.g.,][]{Luhman99,Brandner00,Pontoppidan05,Lahuis06,vanKempen09}. The use of a spectral line map over a small ($\\sim$2$'\\times2'$) region around the protostar is an elegant solution to disentangle the contributions of the different components \\citep{Boogert02}. In particular, the spatial distribution of the HCO$^+$ 4--3 line has proven to be an excellent diagnostic of such `false' embedded sources: in the case of Ophiuchus, about 60\\% of the sources with Class I or flat SED slopes turned out not to be embedded YSOs \\citep{vanKempen09}. With the development of the Atacama Large Millimeter/Submillimeter Array (ALMA) on Chajnantor, Chile, surveys of sources in southern star-forming clouds will be undertaken at high resolution ($\\leq$ 1$''$) at a wide range of frequencies, and are expected to reveal much about the inner structure of protostars and their circumstellar disks. The {\\it Herschel Space Observatory} will also target a large number of low-mass YSOs across the sky. Both sets of observations rely heavily on complementary large aperture ground-based single-dish observations for planning and interpretation. Apart from providing the necessary information about the structure on larger scales, single-dish studies of low-mass star formation in the southern sky will also put the results obtained from studies on the northern hemisphere in perspective. Due to a lack of sub-millimeter telescopes at high dry sites in the southern hemisphere, high frequency data on southern YSOs are still limited. The Swedish ESO Submillimeter Telescope (SEST) operated up to the 345 GHz window, but for only limited periods of the year. As a result, southern clouds, such as Chamaeleon, Corona Autralis, Vela or Lupus, are much less studied than their counterparts in the northern sky, such as Taurus and Perseus, where such observations have been readily available for over a decade \\citep[e.g.,][]{Hogerheijde97,Motte98,Hogerheijde98,Johnstone00,Jorgensen02,Nutter05,Nutter08}. The Atacama Pathfinder EXperiment (APEX) \\citep{Guesten06} \\footnote{This publication is based on data acquired with the Atacama Pathfinder Experiment (APEX) with programs E-77.C-0217, E-77.C-4010 and E-78.C-0576. APEX is a collaboration between the Max-Planck-Institut f\\\"ur Radioastronomie, the European Southern Observatory, and the Onsala Space Observatory.} has opened up access to the atmospheric windows in the 200-1400 $\\mu$m wavelength regime over the entire southern sky. Of the southern star-forming regions not or only poorly visible with northern telescopes two regions have proven especially interesting. The Chamaeleon I cloud ($D$=130 pc, Dec = -77$^\\circ$), observed with {\\it IRAS} and the {\\it Infrared Space Observatory} (ISO) and included in the {\\it Spitzer Space Telescope} guaranteed time and $`$Cores to disks' (c2d) Legacy programs \\citep{Persi00,Evans03,Damjanov07,Luhman08} contains some embedded sources, especially around the Cederblad region \\citep{Persi00,Lehtinen01,Belloche06,Hiramatsu07}. Corona Australis is a nearby star-forming region ($D$=170 pc, see \\citet{Knude98}; Dec = -36$^\\circ$), well-known for the central Coronet cluster near the R CrA star \\citep{Loren79,Taylor84,Wilking86,Brown87}. Large-scale C$^{18}$O 1--0 maps show that it contains about 50 M$_\\odot$ of gas and dust \\citep{Harju93}. Many surveys have been undertaken in the IR \\citep[e.g.,][]{Wilking86, Wilking97,Olofsson99,Nisini05}, but only a few studies mapped this region at submillimeter wavelengths \\citep{Chini03,Nutter05}. Although Corona Australis has only a few protostars with rising infrared SEDs, the cloud does contains some of the most luminous low-mass protostars in the neighborhood of the Sun ($D<$ 200 pc) with luminosities of up to 20 L$_{\\rm{\\odot}}$. In this paper we present observations of submillimeter lines of CO and HCO$^+$ of a sample of 16 embedded sources in the southern sky, with a focus on the Chamaeleon I and Corona Australis clouds, to identify basic parameters such as column density, presence and influence of outflowing material, presence of warm and dense gas and the influence of the immediate surroundings, in preparation for Herschel and ALMA surveys or in-depth high resolution interferometric observations. In section $\\S$ 2 we present the observations, for which the results are given in $\\S$ 3 with a distinction between the clouds and isolated sources. Both the single spectra ($\\S$ 3.1) and maps ($\\S$ 3.2) are presented. In $\\S$ 4 we perform the analysis of the observations, making use of archival submillimeter continuum data, with the final conclusions given in $\\S$ 5. \\placeTableChapterFourOne ", "conclusions": "We present observations of CO, HCO$^+$ and their isotopologues, ranging in excitation from CO 3--2 to CO 7--6 of a sample of southern embedded sources to probe the molecular gas content in both the protostellar envelopes and molecular outflows in preparation for future ALMA and Herschel surveys. The main conclusions are the following: \\begin{itemize} \\item HCO$^+$ 4--3 and CO 4--3 integrated intensities and concentrations, combined with information on the presence of outflows, confirm the presence of warm dense quiescent gas associated with 11 of our 16 sources. \\item RCrA TS 3.5, Cha IRS 6a, Cha INa2, IRAS 07178-4429 and IRAS 13546-3941 are likely not embedded YSOs due to the lack of HCO$^+$ 4--3 emission and/or the lack of central concentration in HCO$^+$ 4--3. \\item The swept-up outflow gas has temperatures of the order of 50--100 K, as measured from the ratios of the CO line profile wings, with the values depending on the adopted ambient densities. These values are comparable to the temperatures predicted by the heating model of \\citet{Hatchell99}, with the exception of IRAS 15398-3359, which may be unusually warm. \\item The outflows of 6 of our truly embedded sources --- Ced 110 IRS 4, CrA IRAS 32, RCrA IRS 7A, HH 100, HH 46,IRAS 15398-3359 --- were characterized using CO spectral maps. The outflows all have exceptionally strong outflow forces, almost two orders of magnitude higher than expected from their luminosities following the relation of \\citet{Bontemps96}. \\item Neither Chamaeleon nor Corona Australis have foreground layers as found in Ophiuchus L~1688. All C$^{18}$O 3--2 spectra can be fitted with single gaussians. \\end{itemize} Future observations using ALMA and Herschel will be able to probe the molecular emission associated with these YSOs at higher spatial resolution and higher frequencies. Comparison with single dish data, both continuum surveys and spectral line mapping, will be essential in analyzing the results obtained with future facilities.\\\\" }, "0910/0910.5220_arXiv.txt": { "abstract": "We generalize the ``renormalized'' perturbation theory (RPT) formalism of \\citet{CrocceScoccimarro2006a} to deal with multiple fluids in the Universe and here we present the complete calculations up to the one-loop level in the RPT. We apply this approach to the problem of following the non-linear evolution of baryon and cold dark matter (CDM) perturbations, evolving from the distinct sets of initial conditions, from the high redshift post-recombination Universe right through to the present day. In current theoretical and numerical models of structure formation, it is standard practice to treat baryons and CDM as an effective single matter fluid -- the so called dark matter only modeling. In this approximation, one uses a weighed sum of late time baryon and CDM transfer functions to set initial mass fluctuations. In this paper we explore whether this approach can be employed for high precision modeling of structure formation. We show that, even if we only follow the linear evolution, there is a large-scale scale-dependent bias between baryons and CDM for the currently favored WMAP5 $\\Lambda$CDM model. This time evolving bias is significant $(>1\\%)$ until the present day, when it is driven towards unity through gravitational relaxation processes. Using the RPT formalism we test this approximation in the non-linear regime. We show that the non-linear CDM power spectrum in the 2-component fluid differs from that obtained from an effective mean-mass 1-component fluid by $\\sim3\\%$ on scales of order $k\\sim0.05\\kMpc$ at $z=10$, and by $\\sim0.5\\%$ at $z=0$. However, for the case of the non-linear evolution of the baryons the situation is worse and we find that the power spectrum is suppressed, relative to the total matter, by $\\sim15\\%$ on scales $k\\sim0.05\\kMpc$ at $z=10$, and by $\\sim3-5\\%$ at $z=0$. Importantly, besides the suppression of the spectrum, the Baryonic Acoustic Oscillation (BAO) features are amplified for baryon and slightly damped for CDM spectra. If we compare the total matter power spectra in the 2- and 1-component fluid approaches, then we find excellent agreement, with deviations being $<0.5\\%$ throughout the evolution. Consequences: high precision modeling of the large-scale distribution of baryons in the Universe can not be achieved through an effective mean-mass 1-component fluid approximation; detection significance of BAO will be amplified in probes that study baryonic matter, relative to probes that study the CDM or total mass only. The CDM distribution can be modeled accurately at late times and the total matter at all times. This is good news for probes that are sensitive to the total mass, such as gravitational weak lensing as existing modeling techniques are good enough. Lastly, we identify an analytic approximation that greatly simplifies the evaluation of the full PT expressions, and it is better than $<1\\%$ over the full range of scales and times considered. ", "introduction": "\\label{sec:intro} In the current paradigm for structure formation in the Universe, it is supposed that there was an initial period of inflation, during which, quantum fluctuations were generated and inflated up to super-horizon scales; producing a near scale-invariant and near Gaussian set of primordial potential fluctuations. At the end of inflation the Universe is reheated, and particles and radiation are synthesized. In this hot early phase, cold thermal relics are also produced, these particles interact gravitationally and possibly through the weak interaction -- dubbed Cold Dark Matter (CDM). Prior to recombination photons are coupled to electrons through Thompson scattering, and in turn electrons to baryonic nuclei through the Coulomb interaction. Baryon fluctuations are pressure supported on scales smaller than the sound horizon scale. Subsequent to recombination photons free-stream out of perturbations and baryons cool into the CDM potential wells. Structure formation then proceeds in a hierarchical way with small objects collapsing first and then merging to form larger objects. Eventually, sufficiently dense gas wells are accumulated and galaxies form. At late times the Universe switches from a decelerating phase of expansion to an accelerated phase. This is attributed to the non-zero constant energy-density of the vacuum. This paradigm has been dubbed: $\\Lambda$CDM \\citep[][]{PeeblesRatra2003,Spergeletal2007,Komatsuetal2009}. The present day energy-density budget for the $\\Lambda$CDM model is distributed into several components, which in units of the critical density are: vacuum energy $\\Omega_{\\Lambda,0}\\approx0.73$, matter $\\Omega_{\\rmm,0}\\approx0.27$, neutrinos $\\Omega_{\\nu,0}\\lesssim 10^{-2}$ and radiation $\\Omega_{\\rr,0}\\approx 5\\times 10^{-4}$. The matter distribution can be subdivided further into contributions from CDM and baryons, with present day values $\\Omega_{\\rc}\\approx0.225$ and $\\Omega_{\\rb}\\approx0.045$. The detailed physics for the evolution of the radiation, CDM, baryon and neutrino fluctuations ($\\delta^{\\rr}, \\delta^{\\rc}, \\delta^{\\rb}, \\delta^{\\nu}$), from the early Universe through to recombination can be obtained by solving the Einstein--Boltzmann equations. These are a set of coupled non-linear partial differential equations, however while the fluctuations are small they may be linearized and solved \\citep[for a review see][]{MaBertschinger1995,SeljakZaldarriaga1996}. At later times these fluctuations enter a phase of non-linear growth. Their evolution during this period can no longer be described accurately using the linear Einstein--Boltzmann theory and must be followed using higher order perturbation theory (hereafter PT) techniques \\cite{Bernardeauetal2002} or more directly through $N$-body simulations \\cite{Bertschinger2001}. \\begin{figure}[t!] \\centering{ \\includegraphics[width=8cm,clip=]{Fig.1.eps}} \\caption{\\small{Baryon bias as a function of inverse spatial scale, where we have defined the baryon bias $b_{\\rb}(k,z)\\equiv \\delta^{\\rb}_{\\mathrm{lin}}(k,z)/\\delta^{\\rc}_{\\mathrm{lin}}(k,z)$, and used \\Eqns{eq:Lin_c}{eq:Lin_b} from \\S\\ref{ssec:application}. Solid through to dashed lines show results for redshifts $z=\\{0, 1.0, 3.0, 5.0, 10.0, 20.0\\}$.}\\label{fig:baryonbias}} \\end{figure} The cross-over from the linear to the non-linear theory is, in general, not straightforward. Most of the theoretical and numerical approaches to this latter task were developed during the previous two decades, during which time the Standard CDM model was favored: essentially Einstein-de Sitter spacetime, $\\Omega_{\\rmm}=1$, with energy-density dominated by CDM $\\gtrsim 97\\%$, and baryons contributing $\\lesssim3\\%$ to the total matter. For this case it is natural to assume that the CDM fluctuations dominate the gravitational potentials at nearly all times, with baryons playing little role in the formation of large-scale structure \\citep[see for example][]{Efstathiouetal1985,ThomasCouchman1992,Couchmanetal1995}. Thus, one simply requires a transfer function for the CDM distribution at some late time, to model the evolution of both components. In the latter part of the 1990s, cosmological tests pointed towards the $\\Lambda$CDM paradigm, and it was recognized that baryons should play some role in shaping the transfer function of matter fluctuations \\citep{MaBertschinger1995,Sugiyama1995,EisensteinHu1998,MeiksinWhitePeacock1999}. However, rather than attempting to follow the CDM and baryons as separate fluids evolving from distinct sets of initial conditions, a simpler approximate scheme was adopted. This scheme is currently standard practice for all studies of structure formation in the Universe \\citep{Pearceetal2001,Teyssier2002,Springel2005,Springel2009}. It can be summarized as follows: \\begin{enumerate} \\item Fix the cosmological model, specifying $\\Omega_{\\rc}$ and $\\Omega_{\\rb}$, and hence the fraction of baryons, $\\fb$. Solve for the evolution of all perturbed species using the linearized coupled Einstein--Boltzmann equations. Obtain transfer functions at $z=0$, where CDM and baryons are fully relaxed: i.e.~$T^{\\rc}(k,z=0)\\approx T^{\\rb}(k,z=0)$, where $T^{\\rc}$ and $T^{\\rb}$ are the transfer functions of the CDM and baryons, respectively. \\item Use the transfer functions to generate the linear matter power spectrum, normalized to the present day: i.e. $P_{\\bar{\\delta}\\bar{\\delta}}(\\bk,z=0)\\approx \\left[(1-\\fb) T^{\\rc}(k,z=0)+\\fb T^{\\rb}(k,z=0)\\right]^2Ak^n \\approx [T^{\\rc}(k,z=0)]^2 Ak^n\\ ,$ where the total matter fluctuation is $\\bar{\\delta}=(1-\\fb)\\delta^{\\rc}+\\fb\\delta^{\\rb}$. \\item Scale this power spectrum back to the initial time $z_i$, using the linear growth factor for the total matter fluctuation $\\bar{\\delta}$, which obtains from solving the equations of motion for a single fluid. \\item Generate the initial CDM density field and assume that the baryons are perfect tracers of the CDM. \\item Evolve this effective CDM+baryon fluid under gravity using full non-linear equations of motion, either as a single fluid for dark matter only, or as a 2-component fluid if hydrodynamics are also to be followed. \\end{enumerate} This effective description for the formation of structure becomes poor as the baryon fraction $\\fb=\\Omega_{\\rb}/\\Omega_{\\rmm}$ approaches $\\sim O(1)$, and for the currently favored WMAP5 cosmology $\\fb\\approx0.17$. In fact, in linear theory the gravitational relaxation between baryonic and dark matter components is only achieved at the level of $<1\\%$ by $z=0$. Moreover, as can be seen in \\fig{fig:baryonbias} there is a non-trivial scale-dependent bias between baryons and dark matter that exists right through to the present day \\footnote{Note that we have defined the baryon bias $b_{\\rb}(k,z)\\equiv\\delta^{\\rb}_{\\mathrm{lin}}(k,z)/ \\delta^{\\rc}_{\\mathrm{lin}}(k,z)$, where these linear fluctuations are given by \\Eqns{eq:Lin_c}{eq:Lin_b} from \\S\\ref{ssec:application}. In this analysis we have neglected any residual effect of the coupling between the photons and the baryons, and have assumed that the baryons are cold at $z=100$, hence $p_{\\rm b}\\propto \\rho T\\sim 0$. Furthermore, this result is sensitive to our choice for the initial $\\delta_i$ and $\\theta_i$, and indeed had we chosen the initial baryon and CDM velocity fields to be proportional to the {\\em total matter} fluctuation, then these effects would be slightly reduced. However, the resultant bias will still remain non-negligible.}. Failing, to take these biases into account may lead to non-negligible systematic errors in studies that attempt to obtain cosmological constraints from Large-Scale Structure (LSS) tests. With the next generation of LSS tests aiming to achieving 1\\% precision in measurements of the matter power spectrum \\cite{DETF2006,ESO2006}, it seems timely to investigate whether such modeling assumptions may impact inferences. Where we expect such systematic effects to play an important role is for studies based on baryonic physics such as: using 21cm emission from neutral hydrogen to trace mass density at the epoch of reionization $z_{\\rm reion} \\approx 10$ \\citep{Zhangetal2004,Ilievetal2006,Furlanettoetal2006,PillepichPorciani2007}; and studies of the Lyman alpha forest to probe the matter power spectrum \\citep{Croftetal1998,McDonaldetal2006}. In this, the first in a series of papers, we test whether this effective procedure can indeed be employed to follow, at high precision, the non-linear growth of structure formation in baryon and CDM fluids, from recombination right through to the present day. We do this by generalizing the matrix based ``renormalized'' perturbation theory (RPT) \\footnote{It appears to us that what may be implemented in the RPT framework is not so much what would usually be understood as the ``perturbative renormalization'' of the theory, but rather a ``resummation'' of the perturbative series, in some specific approximation (the ``high-$k$'' limit). We will continue to use the acronym RPT to refer to the method, but would suggest that it is more properly read as {\\em resummed}, and not {\\em renormalized} perturbation theory.} techniques of \\citet{CrocceScoccimarro2006a,CrocceScoccimarro2006b,CrocceScoccimarro2008} to an arbitrary number of gravitating fluids. In this paper we content our selves with calculating the observables up to the one-loop level in the theory. In a future paper we consider the resummation to arbitrary orders (Somogyi \\& Smith 2009, in preparation). The late-time linear theory evolution of coupled baryon and CDM fluctuations, with sudden late time reionization of the inter-galactic medium was studied by \\citep{Nusser2000,MatarreseMohayaee2002}. Recently, \\citep{ShojiKomatsu2009} explored a similar system using non-linear PT techniques. The work presented by these authors differs from that presented here in a number of fundamental ways. Firstly, these authors have assumed that, on large scales, baryons and CDM fluctuations evolve with the same initial conditions and that the only physical difference between the two fluids arises on small scales, where galaxy formation processes are able to provide some pressure support. They model this with a fixed comoving-scale Jeans criterion. In this work we are not so much interested in following how small-scale baryonic collapse feeds back on the mass distribution, but are more interested in following the very large-scale baryon and CDM fluctuations, initialized with their correct post-recombination spatial distributions, into the non-linear regime. Secondly, their work was developed within standard PT, the problems associated with this formalism are well documented \\citep{Bernardeauetal2002}. Here, we develop the more successful RPT approach \\citep{CrocceScoccimarro2006a}. In passing, we note that a number of additional complimentary approaches and extensions to single-fluid RPT have recently been developed \\citep{MatarresePietroni2007,MatarresePietroni2008,Bernardeauetal2008,Pietroni2008,Matsubara2008}. In addition, there has recently been progress on realistic modeling of massive neutrinos and CDM \\citep{Wong2008,Saitoetal2008,Ichikietal2009,Brandbygeetal2008,BrandbygeHannestad2009a,BrandbygeHannestad2009b,Lesgourguesetal2009}, for simplicity we shall assume that neutrinos play a negligible role. The paper is broken down as follows: In \\Sect{sec:EoM} we set up the theoretical formalism: presenting the coupled equations of motion for the $N$-component fluid. Then we show how, when written in matrix form, the equations may be solved at linear order. We discuss generalized initial conditions, and then discuss the specific case of baryons and CDM. In \\Sect{sec:PT} we develop the non-linear PT expansion for the solutions, and show that they may be constructed to arbitrary order using a set of Feynman rules. In \\Sect{sec:stats} we discuss the two-point statistics of the fields, which we take as our lowest order observables. In \\Sect{sec:OneLoopPower} we give specific details of how we calculate the one-loop corrections in our theory. In \\Sect{sec:results} we present results for various quantities of interest. In \\Sect{sec:approx} we give details of an approximate RPT scheme, which matches the exact calculation to within $<1\\%$, for all times of interest and on all scales where the perturbation theory is valid. Finally, in \\Sect{sec:conclusions} we draw our conclusions. ", "conclusions": "\\label{sec:conclusions} In this paper we have generalized the matrix based RPT formalism of \\citet{CrocceScoccimarro2006a} to the problem of dealing with $N$-component fluids, each of which evolves from a distinct set of initial conditions and contributes to the matter-density and velocity fields. In \\Sect{sec:EoM} we showed that, for $N$-component fluids, the essential structure of the RPT framework remained the same as for the $N=1$ case -- that is, the equations of motion could be solved in exactly the same way. However, the details of the building blocks of the theory did change: i.e.~the vertex, propagator and initial conditions. For $N=2$ the vertex matrix offered no surprises. The linear propagator was more interesting, and we showed that besides the usual growing and decaying modes $\\{D_{+}\\propto \\re^{\\eta}, D^{(1)}_{-}\\propto \\re^{-3\\eta/2}\\}$, there were two additional eigenvalues: a static and decaying mode $\\{D_{\\rm stat}\\propto {\\rm const}, D^{(2)}_{-}\\propto \\re^{-\\eta/2}\\}$. Interestingly, in going to $N>2$ no further time dependence was gained. Unlike the 1-component fluid case, it was no longer possible to choose pure growing modes, unless all of the fluids evolved from identical initial conditions. However, we showed that one can still set up the initial conditions to possess pure large-scale growing modes. All our results were obtained for this specific choice of initial conditions. In \\Sect{sec:PT} it was shown that the equations of motion for the $N$-component fluid case could be solved using a power series solution and that the solutions could be recovered order by order. A graphical representation of the solutions was also described. In \\Sect{sec:stats} we discussed the statistics of the fields: in particular the covariance matrix of the $N$-component fluid perturbations, and the fully non-linear propagator. This latter quantity can be understood to be the statistic that provides information on the memory of the evolved field to the initial conditions. It also possesses a perturbative expansion and we gave expressions up to one-loop order. A diagrammatic representation was also discussed. As for the case of the 1-component fluid RPT, we then showed that for the $N$-component fluid RPT the full non-linear power spectrum also possessed a series expansion, and that the series could be grouped into two sub-series, ``reducible'' and ``mode-coupling''. It was shown that the reducible terms were proportional to the initial power spectrum and non-linear propagators. Again we gave complete details for the series up to the one-loop level in the expansions. A diagrammatic description was also discussed. In \\Sect{sec:OneLoopPower} we developed further the one-loop expressions for the propagator and the mode-coupling contribution to the power spectrum. These expressions were then given for the explicit case of a 2-component fluid. In \\Sect{sec:results} we applied the formalism to the problem of modeling the non-linear evolution of a CDM and baryon fluid in the $\\Lambda$CDM paradigm. This enabled us to test the validity of the approximation of treating CDM and baryons as an effective 1-component fluid evolving from a single set of initial conditions, as is currently standard practice. For the case of CDM, it was shown that the approximation was very good for $z<3$, with the exact and approximate CDM power spectra differing by $<1\\%$. However, for higher redshifts the approximation became progressively worse, it being of the order $\\sim3\\%$ at $z=10$, with the power in the exact theory being amplified and having weaker BAO features than the approximate theory. For the case of baryons the situation was worse: at $z=20$ the exact and approximate spectra differed by $\\sim25\\%$, and at $z=10$ by $\\sim15\\%$. At later times the approximation was still quite poor, and at $z=3$ the errors were of the order $\\sim 5\\%$, and by $z=0$ were still between $\\sim2-3\\%$. The power in the exact theory was suppressed in amplitude but possessed stronger BAO than the corresponding power for the approximate theory. These conclusions remained essentially unaffected, when the so-called approximate 1-component fluid initial conditions were employed for the effective 1-component fluid computation. Lastly, we computed the total non-linear matter power spectrum and found that the exact and approximate theory agreed to within $<0.03\\%$, when the so-called exact 1-component fluid initial conditions were used. Employing the approximate 1-component fluid initial conditions on the other hand leads to an agreement to within $<0.5\\%$ on scales where the perturbation theory was valid $k<0.2\\kMpc$ and for all times of interest. Deviations on the order of $\\sim0.7\\%$ were found on smaller scales with this approximation. \\vspace{0.2cm} Our main conclusions therefore are: \\begin{itemize} \\item For theoretical modeling of the low-redshift CDM distribution, the approximate 1-component fluid treatment provides a good approximation. For higher redshift studies, such as those that probe the epoch of reionization, or attempt to model the first objects that form (stars/haloes), then one must be careful to use the appropriate CDM transfer function for the redshift of interest. Otherwise, systematic errors will be present at the level of several percent. \\item For theoretical modeling of baryons in the Universe, a 1-component fluid treatment leads to significant $>1\\%$ errors at all times. This implies that cosmological probes that are primarily sensitive to the distributions of baryons in the universe, such as: 21 cm radiation from neutral hydrogen, or the Lyman alpha forest absorption lines in high redshift quasars, can not be modeled using dark matter only simulations at high precision. Instead the baryons must be modeled using a 2-component fluid system of CDM and baryons, with the initial distributions being specified by the linear Einstein--Boltzmann solutions. The correct treatment of CDM and baryon initial conditions may also affect how galaxies form. \\item Observational probes for cosmology that are sensitive to the total mass distribution, such as weak gravitational lensing, can be accurately interpreted using an effective single CDM plus baryon fluid. Thus current modeling technology is good enough for high precision work, but it is highly desirable to use the exact prescription when specifying the initial conditions for the effective single fluid. \\end{itemize}" }, "0910/0910.5685_arXiv.txt": { "abstract": "When scientific experiments require transmission of powerful laser or radio beams through the atmosphere the Federal Aviation Administration (FAA) requires that precautions be taken to avoid inadvertent illumination of aircraft. Here we describe a highly reliable system for detecting aircraft entering the vicinity of a laser beam by making use of the Air Traffic Control (ATC) transponders required on most aircraft. This system uses two antennas, both aligned with the laser beam. One antenna has a broad beam and the other has a narrow beam. The ratio of the transponder power received in the narrow beam to that received in the broad beam gives a measure of the angular distance of the aircraft from the axis that is independent of the range or the transmitter power. This ratio is easily measured and can be used to shutter the laser when the aircraft is too close to the beam. Prototype systems operating on astronomical telescopes have produced good results. ", "introduction": "A number of scientific experiments require the transmission of a laser beam through the atmosphere, using an astronomical telescope or its equivalent. In order to avoid hazard to aircraft the FAA requires that one or more observers be stationed outside any telescope that is transmitting a laser beam. These observers close the laser shutter when an aircraft is observed within $25^\\circ$ of the laser beam (as viewed from the telescope). Such experiments include: lunar and satellite laser ranging \\citep{llr,ilrs}; creation of artificial guide stars for active optics \\citep[e.g.,][]{keck-lgs}; and atmospheric remote sensing using lidar \\citep{lidar1,lidar2}. In this paper we discuss part of an aircraft detection system now employed at the Apache Point Observatory (APO) for a lunar ranging experiment called APOLLO \\citep{apollo}. This system could be used not only for the other laser beam experiments mentioned above, but also for protecting aircraft from high-powered radar transmitters such as those used for ionospheric research \\citep[e.g.,][]{ionosphere}. The detection scheme described here is used in conjunction with a complementary infrared camera detection system, together providing robust protection to aircraft. The FAA rules effectively require transponders on all commercial aircraft and most private aircraft \\citep[the exact language can be found in Section 91.215 of the Federal Aviation Regulations:][]{far}. These transponders are interrogated frequently (at 1030~MHz) by the regional ATC radars and also by the airborne Traffic Collision Avoidance System (TCAS). The transponders reply incoherently at $1090\\pm 3$~MHz with a pulse coded response. The response must have vertical electric field polarization, an omni-directional pattern, and transmitted peak power between 70 and 500~W. Various coding schemes convey information about the aircraft. Mode-A and Mode-C responses communicate a temporarily assigned aircraft identity and altitude, respectively. A newer Mode-S encoding flexibly communicates permanent aircraft identity, coordinates, altitude, etc., but still as pulsed transmission at 1090~MHz. The APOLLO laser is never used at elevations less than $15^\\circ$ and this elevation restriction is typical of other laser and radar transmitters. Thus for practical altitudes of $< 13$~km the aircraft range will not exceed $\\sim$50~km, and the received power is very high by modern communications standards ($> -69$~dBm) so that it may be easily detected with a total power receiver. However the received power is highly variable because both the range and the transmitted power are variable. The design requirement is a highly reliable method of detecting when an aircraft transponder is within about $15^\\circ$ of the telescope beam. This $15^\\circ$ specification---differing from the 25$^\\circ$ angle used by human spotters---is set by the expected angular rate of aircraft, transponder interrogation frequency, and the desire to avoid excessive triggers when pointing the beam as low as $15^\\circ$ above the horizon. The general concept is to use two antennas aligned with the optical axis of the telescope, one with a beam width (full-width at half-power) of about $30^\\circ$ and the other with a beam width of about $90^\\circ$, as shown in Figure~\\ref{fig:gains}. The ratio of the power received by the narrow beam antenna to that received by the broad beam antenna depends only on the angular position of the transponder with respect to the beam axis. In particular it does not depend on the distance, transmitted power, or polarization mismatch. \\begin{figure} \\begin{center}\\includegraphics[width=89mm, angle=0]{fig1.pdf} \\end{center} \\caption{Gain of the broad-beam and narrow-beam antennas as a function of the angular separation of the transponder from the axis. The solid lines are H-plane cuts through the gains of the antennas actually used and the dashed lines are E-plane cuts. If the antennas are pointed to zero elevation, the E plane is the vertical plane and the H plane is horizontal. The power received is proportional to the gain. \\label{fig:gains}} \\end{figure} The APO telescope, like many others used in this type of experiment, is on an altazimuth mount with a secondary mirror more than 80~cm in diameter centered on the optical axis. The radio antennas can be mounted facing the sky on the secondary support structure and aligned with the optical axis without interfering with the optical beam, and the polarization will remain vertical as the telescope is moved. This geometry is also typical of the radar antennas used for ionospheric research although the secondary reflector is much larger in these cases. For telescopes on an equatorial mount the position angle of the linear polarization changes with hour angle. This variation can be easily accommodated by changing to antennas that are sensitive to circular polarization. Simple patch antennas are well-suited to this application because the narrow bandwidth of a simple patch is an advantage when the signal also has a narrow bandwidth. Patches are also very robust mechanically---a significant advantage in this application. The polarization of a patch can be changed from linear to circular simply by moving the feed point. A single patch is suitable for the broad beam antenna and the narrow beam antenna can be made with an array of patches. The ratio of the power in the array to the power in a single patch will depend only on the array factor, which is easily calculated. The element spacing of the array can be adjusted to optimize the beam width and the sidelobe levels. The system design consists primarily of: impedance matching a simple patch antenna at 1090~MHz; adjusting the array configuration to obtain a suitable beam width and adequately low sidelobes; designing total power detectors for the two channels; development of signal processing electronics; and devising a reliable calibration system. A block diagram of the analog signal flow is shown in Figure~\\ref{fig:block_diagram}. Here one can see that we have used the center element of the array both as an array element and as the broad beam element by splitting the signal with a power divider. To compensate for that power division, the other array elements must have -3~dB attenuators before the array summer. \\begin{figure} \\begin{center}\\includegraphics[width=170mm, angle=0]{fig2.pdf} \\end{center} \\caption{Block diagram of the analog signal flow. The patches are shown with the E field vertical. The azimuthal angle is defined with respect to the horizontal. The array is drawn approximately to scale. The center patch is used both as an array element and as the broad beam element. \\label{fig:block_diagram}} \\end{figure} ", "conclusions": "" }, "0910/0910.2546_arXiv.txt": { "abstract": "Most stars are members of binaries, and the evolution of a star in a close binary system differs from that of an ioslated star due to the proximity of its companion star. The components in a binary system interact in many ways and binary evolution leads to the formation of many peculiar stars, including blue stragglers and hot subdwarfs. We will discuss binary evolution and the formation of blue stragglers and hot subdwarfs, and show that those hot objects are important in the study of evolutionary population synthesis (EPS), and conclude that binary interactions should be included in the study of EPS. Indeed, binary interactions make a stellar population younger (hotter), and the far-ultraviolet (UV) excess in elliptical galaxies is shown to be most likely resulted from binary interactions. This has major implications for understanding the evolution of the far-UV excess and elliptical galaxies in general. In particular, it implies that the far-UV excess is not a sign of age, as had been postulated prviously and predicts that it should not be strongly dependent on the metallicity of the population, but exists universally from dwarf ellipticals to giant ellipticals. ", "introduction": "Evolutionary population synthesis (EPS) has experienced a rapid progress since the early 90's and provides the most robust approach in studying stellar populations of galaxies. In most of the current EPS models, binary evolution has been ignored. However, most stars are members of binaries, and binary evolution leads to the formation of many peculiar objects, such as blue stagglers and hot subdwarfs, which are hot, long-lived and still very luminous. Those objects contribute very much to the spectral energy distribution (SED) at short wavelength for an old stellar population, and make the population look hotter and younger. Such a ``cosmetic'' effect has been successfully included by EPS models of \\cite[Zhang et al.\\ (2004)]{zha04}, \\cite[Han, Podsiadlowski \\& Lynas-Gray (2007)]{han07}, \\cite[Chen \\& Han (2009)]{che09}. Some puzzles in EPS have been solved with the inclusiong of binaries and binary evoltuion is an important ingredient in EPS and also a subject of this symposium. ", "conclusions": "" }, "0910/0910.5562.txt": { "abstract": "We monitored the BL Lac object S5 0716+714 in the optical band during October 2008, December 2008 and February 2009 with a best temporal resolution of about 5 minutes in the \\emph{BVRI} bands. Four fast flares were observed with amplitudes ranging from 0.3 to 0.75 mag. The source remained active during the whole monitoring campaign, showing microvariability in all days except for one. The overall variability amplitudes are $\\Delta$\\emph{B} $\\sim$ 0$^{m}$.89, $\\Delta$\\emph{V} $\\sim$ 0$^{m}$.80, $\\Delta$\\emph{R} $\\sim$ 0$^{m}$.73 and $\\Delta$\\emph{I} $\\sim$ 0$^{m}$.51. Typical timescales of microvariability range from 2 to 8 hours. The overall \\emph{V} - \\emph{R} color index ranges from 0.37 to 0.59. Strong bluer-when-brighter chromatism was found on internight timescales. However, different spectral behavior was found on intranight timescales. A possible time lag of $\\sim$ 11 mins between \\emph{B} and \\emph{I} bands was found on one night. The shock-in-jet model and geometric effects can be applied to explain the source's intranight behavior. ", "introduction": "Blazars represent an extreme subclass of active galactic nuclei. They are characterized by rapid and strong variability throughout the entire electromagnetic wavebands, high and variable polarization($>$3\\%) from radio to optical wavelengths. In the unified model of AGNs, blazars make an angle of less than $10\\,^{\\circ}$ from the line of sight (Urry \\& Padovani 1995). For low-energy peaked (or radio-selected) blazars the continuum from radio through the UV or soft X-rays is mainly contributed by synchrotron radiation, while a second hump in the spectrum at higher energies usually peaks in the GeV $\\gamma$-ray band.\u00a1\u00b1\\ Blazars exhibit variability on different timescales, from years down to hours or less (See Fan et al. 2005). Understanding blazar variability is one of the major issues of active galactic nuclei studies. There may be different behavior on different timescales. Periodicity may be found on the long term, as is the case of OJ 287 (Sillanpaa et al. 1988) and other blazars (Fan et al. 2002, 2007). The periodicity of OJ 287 can be explained by a binary black hole model, with the secondary black hole passing through the accretion disk of the primary black hole twice per revolution (Sillanpaa et. al 1988; Lehto \\& Valtonen 1996). On shorter timescales, 3C 66A was claimed to have an optical period of $\\sim$ 65 days during its bright state (Lainela et al. 1999), while Mkn 501 may have displayed a period of 23 days in high energy data (Osone 2006). Recently, analyses of X-ray data have yielded excellent evidence for a quasi-period of about an hour for 3C 273 (Espaillat et al. 2008) and very good evidence for near periods of $\\sim$ 16 days for AO 0235+164 and $\\sim$ 420 days for 1ES 2321+419 (Rani et al. 2009). This may be due to a shock wave moving along a helical path in the relativistic jet (e.g., Marscher 1996) or the unstability in the accretion disk (Fan et al. 2008) for Mkn 501. Different viewing angles may also result in different level of brightness, with the source being dimmer at larger viewing angles. BL Lacertae shows variability on both intranight and internight timescales with different spectral behaviors. It shows an intranight strongly chromatic trend and an internight midly chromatic trend, indicating two different components operating in the engine (Villata et al. 2004; Hu et al. 2006). Understanding the shortest variability timescale is of special importance. Brightness changes of up to a few tenths of a magnitude over the course of a night or less is known as intra-night optical variability (INOV) (Wagner \\& Witzel 1995) or the so-called microvariability. It can bring new insight into the understanding of the nature of blazars, probing into the innermost structure down to microparsecs, putting constraints on the size of the source and its physical environments. Microvariability was first discovered in the sixties by Matthews and Sandage (1963), who found 3C 48 changed $0^{m}$.04 in the \\emph{V} band in 15 minutes. But their results were not taken seriously and were considered due to instrumental errors. However, with the development of CCD, microvariability was confirmed to be the intrinsic nature of active galactic nuclei, especially for blazars (e.g., Miller et al. 1989). From a complete sample of BL Lac objects (Stickel et al. 1991), Heidt and Wagner (1996) found 80\\% exhibited microvariability, proving it to be the nature of BL Lac objects. Romero et al. (1999) detected microvariability in 60\\% of their selected sample of 23 southern AGNs. Microvariability became extensively observed since the eighties. Up to now, the reasons for it are still unclear. Many different models have been proposed to explain this phenomenon. It conceivably may be due to eclipses if there is a binary black hole system (Xie et al. 2002). The interaction of relativistic shock waves and inhomogeneous jet can also cause microvariability in the jet (Maraschi et al. 1989; Qian et al. 1991; Marscher 1992; Marscher 1996). Other models like instabilities and perturbations in the accretion disk can explain some of the microvariability phenomena (e.g., Mangalam \\& Wiita 1993; Wiita 1996; Fan et al. 2008). In order to understand the radiation mechanism and the structure of the radiating region, long-term and multiwavelength observations are needed. The BL Lac object S5 0716+714 is one the brightest BL Lac objects noted for its microvariability. Its high duty cycle means that it is always in an active state (Wagner \\& Witzel 1995). It has been the target of many monitoring programs (e.g., Wagner et al. 1996; Qian et al. 2001; Raiteri et al. 2003). Montagni et al. (2006) reported the fastest variability rate of 0.1-0.12mag/hr. Five major outbursts have been observed so far, occurring at the beginning of 1995, in late 1997, in the fall of 2001, in March 2004 and at the beginning of 2007 (Raiteri et al. 2003; Foschini et al. 2006; Gupta et al. 2008). These five outbursts indicate a long-term variability timescale of $\\sim$ 3.0$\\pm$0.3 years (e.g., Raiteri et al. 2003). Correlated radio/optical variability has been observed for the source. In a 4-week monitoring program, Quirrenbach et al. (1991) discovered a period change of 1 day to 7 days in both radio and optical bands. Heidt \\& Wagner (1996) reported a period of 4 days in the optical band while Qian, Tao \\& Fan (2002) derived a 10-day period from their 5.3 yr optical monitoring. Recently Gupta et al. (2009) have found good evidence for quasi-periods in five of the 20 best quality nightly data sets of Montagni et al. (2006); these ranged from $\\sim$ 23 to $\\sim$ 75 minutes. The spectral change of S5 0716+714 has been observed extensively (e.g., Ghisellini et al. 1997; Raiteri et al. 2003; Villata et al. 2004; Gu et al. 2006; Wu et al. 2005). Different behaviors have been reported. Raiteri et al. (2003) reported different chromatism in different timescales in their 8-year monitoring program. Sometimes the source was bluer when brighter, sometimes the opposite, sometimes no spectral change was seen despite changes in brightness. Ghisellini et al. (1997) and Wu et al. (2005) found the source exhibited a bluer-when-brighter chromatism when it was in fast flares. However, Stalin et al. (2006) found even if the source was in fast flare, there was no color change with brightness. In this paper we concentrate on the microvariability and spectral changes of S5 0716+714. We monitored the source from October 25 to 30 2008, December 23 to 29 2008 and February 3 to 10 2009. The temporal resolution was around 5 to 8 minutes in the four optical bands (\\emph{BVRI}). Because of the high temporal resolution, we can provide high quality data with accurate results. The paper is organized as follows: Section 2 describes observations and data reduction procedures. Section 3 presents the results. Discussions are reported in Section 4. A summary is given is Section 5. ", "conclusions": "The microvariability of S5 0716 + 714 has been observed by many authors. Some ultra-rapid fluctuations on timescale $\\leq$ 1.0 hour were reported. Qian et al. (2002) recorded a brightness increase of $\\Delta$\\emph{V} $\\sim$ 0.78 mag in 9 mins in their 5.3-yr monitoring programme. Xie et al. (2004) reported $\\Delta$\\emph{B} $\\sim$ 0.55 mag on a timescale of 36 mins. On longer timescales, Gu et al. (2006) found the magnitude change of $\\Delta$\\emph{V} = 0.28 mag in $\\sim$ 5 hours. In our monitoring programme, no ultra-rapid fluctuations were detected. The shortest timescale was $\\sim$ 2 hours, corresponding to $\\Delta$R $\\sim$ 0.046 mag on JD 2454826 while the longest timescale were $\\sim$ 8 hours, corresponding to $\\Delta$R $\\sim$ 0.245 mag, which happened on JD 2454871 - JD 2454872. In our whole monitoring campaign, the timescales are generally a few hours with amplitude of $\\sim$ 0.04 - 0.28 mag. Time lags between different optical bands have also been reported. Stalin et al. (2006) found possible time lags of $\\sim$ 6 and $\\sim$ 13 mins between the \\emph{V} and \\emph{R }bands in their 2 nights of observations of the source. Qian et al. (2000) reported similar results in which they found a time lag of $\\sim$ 6 mins between the \\emph{V} and \\emph{I} bands. From densely sampling data, Villata et al. (2000) presented a time lag with a strict upper limit of $\\sim$ 10 mins between the \\emph{B} and \\emph{I} band. In this work, we tried to find out the time lag between \\emph{B} band and \\emph{I} band using the discrete cross-correlation function, DCF, suggested by Edelson \\& Krolik (1988) for unequally spaced data. It was performed only on the day with high-quality data (error $\\sim$ 0.003 and $\\sim$ 0.005 for \\emph{B} and \\emph{I} band respectively) and dense sampling rate (temporal resolution $\\sim$ 5 minutes), which is JD 2454826, the day with with a hint of periodicity in its light curve. Fig. 16 presents the results of the calculations. The dashed line indicates the centroid which was calculated using the method proposed by Peterson (2001) and we found the barycenter using data points located near the peak value, DCF$_{peak}$, specifically, those greater than 0.8DCF$_{peak}$. The expression is: \\begin{equation} DCF_{centroid} ={\\frac{\\sum \\tau DCF(\\tau)}{\\sum DCF(\\tau)}} \\end{equation} We found a time lag of $\\sim$ 11 minutes, with the \\emph{B} band leading \\emph{I} band. However, given the sampling rate of $\\sim$ 5 minutes and the negligible difference between the DCF values at -20 minutes and 0 minutes, this cannot be considered a convincing measurement of a lag. This result is consistent with Villata et al. (2000), who presented a strict upper limit of $\\sim$ 10 minutes to a possible delay between \\emph{B} - and \\emph{I} - band variations using high-quality, densely sampled data on a single night. This result is expected in the shock-in-jet model but not in most disc-based variability models (e.g. Wiita 2006). In order to quantitatively analyze possible periods, we performed structure function (SF) on JD 2454826, the only day with a lightcurve that shows a hint of a period. SF, discussed fully by Simonetti el al. (1985), is a common tool to search for periodicities and timescales of variation . It identified a timescale of 259 mins, 287 mins, 280 mins and 263 mins and a period of 482 mins, 497 mins, 498 mins and 491 mins were found for \\emph{B}, \\emph{V}, \\emph{R} and \\emph{I} band respectively. All these results are consistent with visual inspection. The results are shown in Fig. 17. Of course, since only one ``cycle'' is present that night, the presence of a period is no more than speculative; such a possible period does not appear to extend to the previous night, for which there is also data. Only observations made with multiple ground based telescopes at different longitudes (or space based telescopes) can convincingly find intranight periods that substantially exceed $\\sim$ 1 hr. The spectral variability of blazars has been investigated by many authors (e.g., Ghisellini et al. 1997; Fan \\& Lin 1999; Romero et al. 2000; Raiteri et al. 2003; Villata et al. 2000, 2004). Raiteri et al. (2003) found different spectral behavior for the S5 0716 + 714 in short timescales. Sometimes the source was bluer when brighter, sometimes the opposite and sometimes no spectral change was seen. Stalin et al. (2006) found no clear evidence of color variation with brightness in either their internight or intranight monitoring of the source for a fortnight. Ghisellini et al. (1997) and Wu et al. (2005) reported a BWB trend during fast flares but this trend is insensitive in the long-term. However, Wu et al. (2007) noticed that the source is bluer when brighter on both intranight and internight timescales but this trend was not present in the long-term data. In our monitoring campaign, the source also showed different spectral change behavior. On the long term, the source showed a bluer-when-brighter behavior. Unlike previous studies which reported consistent trends on spectral behavior on intranight timescales (e.g. Ghisellini et al. 1997, Stalin et al. 2006, Wu et al. 2005), no consistent spectral behavior is found among all those nights displaying microvariability in our present study. The source displayed BWB chromatism when it was either bright or dim, or when showing large or small variability amplitudes. Achromatism was also found when the source was in the same conditions, suggesting different mechanisms for microvariability. The BWB behavior is consistent with shock-in-jet model (Wagner et al. 1995). In this model, shocks form at the base of the jet and propagate downstream, accelerating electrons and compressing magnetic fields and resulting in the observed variability. The model predicts a BWB phenomenon and an irregular lightcurve, as observed in our cases showing this phenomenon. In the case of JD 2454825 and JD 2454826, a symmetric lightcurve was observed for the former case while a periodic lightcurve for the latter. Such regular lightcurves are rarely found. Wu et al. (2005) reported a single ``period'' of a sine-like light curve on intranight timescales for two nights during their observations of S5 0716+714. Using more sophisticated techniques Gutpa et al. (2009) found multiple oscillations to be present during 5 nights out of the 20 highest quality light curves of a total sample of 102 nights of data by Montagni et al. (2006). Unlike ours, their lightcurves are sinelike with smooth turns while ours are sawtooth-like with sharp turnoffs. The symmetric lightcurves on JD 2454825 showed a correlation between color and magnitude, with the correlation coefficient r = 0.616. So the variation may still be due to intrinsic reasons. However, for the periodic lightcurve on JD 2454826, no strong correlation between color and magnitude is found, with the correlation coefficient r = 0.150. Such achromatic lightcurves may be produced by geometric effects like microlensing or a lighthouse effect produced by different amounts of Doppler boosting induced by a helical structure to the jet (e.g. Wagner \\& Witzel 1995). Microlensing predicts a strictly symmetric lightcurve, which cannot explain the concave shape of the second halves of the light curves in our case. Therefore a lighthouse effect is the most probable mechanism for explaining periodic components in blazar lightcurves as this type of variation is likely to be achromatic (e.g., Camenzind \\& Krockenberger 1992; Gopal-Krishna \\& Wiita 1992)." }, "0910/0910.0225_arXiv.txt": { "abstract": "{Plasma processes close to supernova remnant shocks result in the amplification of magnetic fields and in the acceleration of electrons, injecting them into the diffusive acceleration mechanism.} {The acceleration of electrons and the magnetic field amplification by the collision of two plasma clouds, each consisting of electrons and ions, at a speed of 0.5c is investigated. A quasi-parallel guiding magnetic field, a cloud density ratio of 10 and a plasma temperature of 25 keV are considered.} {A relativistic and electromagnetic particle-in-cell simulation models the plasma in two spatial dimensions employing an ion-to-electron mass ratio of 400.} {A quasi-planar shock forms at the front of the dense plasma cloud. It is mediated by a circularly left-hand polarized electromagnetic wave with an electric field component along the guiding magnetic field. Its propagation direction is close to that of the guiding field and orthogonal to the collision boundary. It has a frequency too low to be determined during the simulation time and a wavelength that equals several times the ion inertial length. These properties would be indicative of a dispersive Alfv\\'en wave close to the ion cyclotron resonance frequency of the left-handed mode, known as the ion whistler, provided that the frequency is appropriate. However, it moves with the super-alfv\\'enic plasma collision speed, suggesting that it is an Alfv\\'en precursor or a nonlinear MHD wave such as a Short Large-Amplitude Magnetic Structure (SLAMS). The growth of the magnetic amplitude of this wave to values well in excess of those of the quasi-parallel guiding field and of the filamentation modes results in a quasi-perpendicular shock. We present evidence for the instability of this mode to a four wave interaction. The waves developing upstream of the dense cloud give rise to electron acceleration ahead of the collision boundary. Energy equipartition between the ions and the electrons is established at the shock and the electrons are accelerated to relativistic speeds.} {The magnetic fields in the foreshock of supernova remnant shocks can be amplified substantially and electrons can be injected into the diffusive acceleration, if strongly magnetised plasma subshells are present in the foreshock, with velocities an order of magnitude faster than the main shell.} ", "introduction": "Supernova remnants (SNRs) emanate energetic electromagnetic radiation, which demonstrates the acceleration of electrons to ultrarelativistic speeds \\citep{RelEl1,Radiation} and the generation or amplification of magnetic fields \\citep{MagAmp1,MagAmp2}. The likely origin of the accelerated electrons and of the strong magnetic fields is the SNR shock \\citep{Marco1,Marco2}. The nonrelativistic expansion speed of the main SNR blast shell \\citep{Shockspeed1,Shockspeed2} and the weak magnetic field of the ambient medium \\citep{MagAmp2}, into which this shell is expanding, are obstacles to the magnetic field amplification by plasma instabilities and to the electron acceleration out of the thermal plasma distribution to moderately relativistic energies. Such an acceleration is needed for their injection \\citep{Injection1,Injection2,Injection3,Gyrosurf} into the diffusive shock acceleration process (See \\citet{Diffus}) so that they can cross the shock transition layer repeatedly. Electrostatic instabilities dominate for nonrelativistic flows in unmagnetized plasmas \\citep{Bret,BRETAPJ} and they can neither accelerate the electrons to highly relativistic speeds \\citep{Sircombe} nor amplify the magnetic fields. The electrons could be accelerated by plasma based charged particle accelerators \\citep{PlAc}, by electron surfing acceleration \\citep{Surfing1,Surfing2,Surfing3}, double layers \\citep{DL2,DL1} or by processes that exploit a velocity shear in the plasma outflow \\citep{Shear}. It has, however, not yet been demonstrated with multi-dimensional and self-consistent simulations that these mechanisms can indeed achieve the required electron acceleration and magnetic field amplification. The non-relativistic shocks, which are found between the main SNR blast shell and the ambient medium, can also probably not transfer significant energy from the ions to the electrons and accelerate the latter to relativistic speeds \\citep{Shock4,Shock6,Shock5,Shock3,Sorasio,Shock2,Shock7}. A viable acceleration mechanism may develop ahead of the main SNR shock, if we find subshells that outrun the main blast shell. These subshells may move faster than the typical peak speed of 0.2c of the main shell. An expansion speed as high as 0.9c may have been observed for a subshell ejected by the supernova SN1998bw \\citep{Shockspeed1}. Most supernovae are less violent and their subshells are probably slower. The density of the subshell plasma is well below that of the main blast shell and its dynamics will be influenced to a larger extent by the upstream magnetic field than the dynamics of the latter. This is true in particular, if the upstream magnetic field has been pre-amplified by the cosmic rays \\citep{Winske,MagAmp3,MagAmp4,Pohl,Riq}. A fast magnetized shock would form in the foreshock of the main SNR shock that can result in a stronger magnetic field amplification and electron acceleration. We examine with a particle-in-cell (PIC) simulation the formation of a shock in a plasma, in which a strong guiding magnetic field is quasi-parallel to the plasma expansion direction. Whistler waves, which become Alfv\\'en waves at low frequencies, occuring at such shocks can be efficient electron accelerators \\citep{Oblique3,Injection3,Gyrosurf,ANOTHERSLAM,Ken}. Whistlers and Alfv\\'en waves are circulary polarized if they propagate parallel to the guiding magnetic field. We briefly summarize their properties and focus on the low-frequency modes with a left-hand polarization. These modes are qualitatively similar to the quasi-parallel propagating ones we consider here. A more thorough description of the dispersion relation of two-fluid waves and the shift of the resonance frequencies for quasi-parallel propagation can be found elsewhere \\citep{Treumann}. The dominant waves, which we will observe, grow in a plasma with an electron gyrofrequency that exceeds the plasma frequency. As we increase the frequency in such a strongly magnetized plasma towards the ion cyclotron frequency, the Alfv\\'en waves with a left-hand circular polarization become dispersive. These waves resonate with the ions and their frequencies remain below the ion cyclotron frequency. Alfv\\'en modes with a left-hand circular polarization just below the ion cyclotron frequency are called ion whistlers. Whistlers are predominantly electromagnetic for small propagation angles relative to the guiding magnetic field, as it has been discussed for high-frequency ones by \\citep{Tokar}, and if they have low wavenumbers. Ion whistlers or dispersive Alfv\\'en waves develop a field-aligned electric field component close to the resonance frequency, by which they can interact nonlinearly with the plasma particles and accelerate them \\citep{Ken}. Any wave growth will furthermore result in an increasing energy density of the magnetic field. Alfv\\'en waves and other magnetohydrodynamic (MHD) waves can grow to amplitudes, at which they start to interact nonlinearly with the plasma \\citep{STASI}. Short Large Amplitude Magnetic Field Structures (SLAMS) are nonlinear MHD waves occuring at quasi-parallel shocks and may be relevant for our simulation. The SLAMS can be efficient electron accelerators \\citep{CME2}. Their magnetic amplitude reaches several times that of the background field and they can propagate with a super-Alfv\\'enic speed, because they convect with the ions \\citep{SLAMSPEED,SLAMS}. SLAMS have also been observed in simulations \\citep{Scholer}. The acceleration of electrons by sub-structures of quasi-parallel shocks, which may be SLAMS, has been observed in the solar corona \\citep{CME}. The absence of self-consistent kinetic models of oblique shocks implies that they can currently be studied only numerically with particle-in-cell (PIC) \\citep{Code2,Code1} or with Vlasov simulations \\citep{Arber,Sircombe}. The pioneering PIC simulations of plasma slabs, which collide with a speed of 0.9c and in the presence of an oblique magnetic field \\citep{Oblique1,Oblique2}, have evidenced the formation of a shock that accelerated the electrons to ultrarelativistic speeds and amplified the magnetic field. A more recent PIC simulation study \\citep{Shock1} has shown, that the shock formation is triggered by an energetic electromagnetic structure (EES). The simulation could demonstrate that an approximate equipartition of the ion, electron and magnetic energy densities is established. However, these simulations could only resolve one spatial dimension due to computer constraints, which is not necessarily realistic for mildly relativistic collision speeds. We perform here a case study with initial conditions, which are similar to those employed by \\citet{Shock1}. We reduce the collision speed to 0.5c and lower the temperature. The plasma cloud representing the subshell is ten times denser than the plasma cloud that represents the ambient plasma (interstellar medium), into which the shell expands. A guiding magnetic field is quasi-parallel to the plasma flow velocity vector and it results in an electron gyrofrequency that equals the electron plasma frequency of the dense cloud. These bulk plasma parameters have been selected with the intention to enforce a planar shock front \\citep{Nish,Mag,Shock1}, by which we can model the shock in one- or two-dimensions in space. The reduced ion-to-electron mass ratio we use allows us to model this collision in form of a 2(1/2)D particle-in-cell (PIC) simulation, which resolves two spatial and three momentum dimensions. Our results are summarized as follows: We find higher-dimensional structures (density filaments), which initially form at the front of the tenuous plasma cloud and expand in time. The front of the dense cloud, which turns out to be the most relevant structure, remains planar and its filamentation is delayed but not suppressed by the guiding magnetic field and the high temperature. An EES grows ahead of the dense plasma cloud before it has become filamentary and the magnetic amplitude of the EES reaches a value several times the one of the initial guiding magnetic field. The EES is pushed by the dense cloud and its phase speed in the rest frame of the tenuous cloud is comparable to the cloud collision speed or twice the Alfv\\'en speed in the tenuous cloud. The front of the EES expands at an even higher speed. Its high speed and amplitude may imply that the EES is a SLAMS. The front of the dense cloud and, consequently, the EES are slowed down by the electromagnetic wave-particle interaction. It is thus not possible to define a rest frame moving with a constant speed, in which the EES is stationary. This would be necessary to measure the frequency of the EES accurately. However, the amplitude distribution of the EES suggests that its oscillation frequency is below the ion cyclotron frequency. Electromagnetic waves are destabilized by the EES ahead of the cloud overlap layer through what we think is a four-wave interaction \\citep{Instability}. The electrons are accelerated in the combined wave fields of these waves and in the forming shock to highly relativistic speeds. The simulation shows though that the strongest electron acceleration occurs at the position, where the shock-reflected ion beam is forming. The plasma collision results in a substantial electron acceleration and also in an amplification of the magnetic energy density by one order of magnitude within the EES and the forming shock. Both values are probably limited by the reduced ion mass of 400 electron masses, which we must employ. Radiative processes, which are not resolved by the PIC code, will at this stage start to influence the shock evolution \\citep{Schl1,Schl2} and we stop the simulation. This paper is structured as follows. Section 2 discusses the particle-in-cell simulation method and the initial conditions. Section 3 presents our results, which are discussed in more detail in Section 4. ", "conclusions": "We have described in this paper the collision of two plasma clouds at the speed c/2. The ion to electron mass ratio of 400 has allowed us to model the collision in two spatial dimensions until the shock forms. Then we had to stop the simulation. The acceleration of electrons to speeds $\\sim c$ and the rapid expansion speed of the energetic electromagnetic structure (EES) imply, that both will quickly reach the boundaries; hence our periodic boundary conditions become invalid. Open boundaries would allow the electrons and the wave energy to flow out of the system. However, the instabilities driven by these beams \\citep{Martins} and by the EES are important for the upstream dynamics and the latter will be adversely affected by open boundaries. The filamentary structures can form and merge in the 2D geometry we consider here up to the instant when the magnetic repulsion of two filaments with oppositely directed current enforces their spatial separation \\citep{Davidson,Lee}, at least in the absence of an oblique magnetic field. Then no further mergers occur since only one dimension is available orthogonal to the flow velocity vector. A realistic 3D PIC simulation would allow the filaments to move around each other and merge with other filaments of equal polarity \\citep{Lee}. A 3D PIC simulation is, however, currently impossible for our case study involving ions, because of the computational cost involved in resolving ion and electron scales. Simulations in three spatial dimensions are now feasible for the case of leptonic shocks \\citep{Nishi2}. The densities of the clouds differ by a factor of 10 and the collision is asymmetric, as in the simulation by \\citet{Oblique1,Oblique2,Sorasio}. Initially the magnetic field is uniform and quasi-parallel to the flow velocity vector. Its significant strength together with the high plasma temperature of 25 keV and the unequal cloud densities reduce the growth rate of the filamentary instabilities \\citep{Mag,Bellido}. \\citet{Shock1} have previously probed the higher $\\gamma$ regime appropriate for the internal shocks of gamma-ray bursts. Here we consider the mildly relativistic regime, with a collision speed $0.5c$ between both clouds. Such a speed might be realistic for a plasma subshell outrunning the main SNR shock. Such subshells can reach relativistic flow speeds for particularly violent SNR explosions \\citep{Shockspeed1}. In our initial conditions, the magnetic energy surpasses the thermal energy of the dense plasma slab by a factor of 5. The plasma flow implies, however, that the box-averaged plasma kinetic energy density exceeds the magnetic energy density by an order of magnitude. We summarize several aspects of our results. \\subsection{Effects due to initial conditions} Our initial conditions have resulted in the formation of planar wave and plasma structures, the most important one being the EES. We think that the EES grew out of a localized seed magnetic field pulse driven by the spatial gradient of the convection electric field. The plasma upstream of the dense cloud is destabilized by this electromagnetic structure and the EES expands at the speed $0.87c$ in the reference frame of the tenuous cloud. The energy for its growth and expansion is provided by the kinetic energy of the upstream medium, which moves with respect to the EES. The shock speed $\\la v_c$ then implies that we have a coherent magnetic layer that expands its width at a speed of at least $(0.87c-v_c)/(1-0.87 v_c / c) \\approx 0.65c$, measured in the reference frame of the tenuous cloud. It covered about 80 ion skin depths at the end of the simulation, showing no signs of a slowdown. The EES is a consequence of our initial conditions and the growth of the seed magnetic field amplitude could probably be reduced but not suppressed by a smoother change of the convection electric field, which can be achieved by a gradual change of the plasma convection speed \\citep{Oblique1,Oblique2}. However, the seed magnetic field could be provided also by waves with a short wavelength, e.g. whistlers, and it is thus not unphysical. Structures with strong magnetic fields, similar to the EES and known as SLAMS, are frequently observed close to quasi-parallel shocks in the solar system plasma and they can accelerate electrons to high energies. They are thus potentially important also for SNR shock physics. Our initial conditions provide a possibility to let nonlinear MHD waves grow out of a simple simulation setup for a further study. The EES is moving with the ions of the dense cloud and it modulates the ions and electrons of the tenuous cloud, thereby gaining energy. Its growth probably requires an asymmetric plasma collision. \\subsection{Shock formation} In this paper we modelled the formation of the shock from the initial collision of two plasma clouds. The signatures of the shock are evident, including visible thermal broadening behind the shock and a dense shock ramp. While filamentation structures form ahead of and behind the shock, we note that the structure is basically planar in the critical foreshock area, where electron acceleration is expected to occur. This means that one-dimensional simulations will be relevant in this region, allowing much higher resolution, increased particle number (resulting in lower particle noise and a better phase space resolution) and a higher ion-electron mass ratio, than is currently found in two and three-dimensional simulations. It is evident from the simulation that the filamentation is not fully suppressed by the guiding magnetic field and by the high plasma temperature. Its amplitude has been set such that it should suppress the electron filamentation if the plasma would be spatially uniform \\citep{Mag}. This amplitude is apparently insufficient to suppress the slower filamentation of the ion beams and we could even see evidence for an electron beam filamentation just behind the front. The front of the dense plasma cloud maintains its planarity throughout the simulation, but even here the density was non-uniform along the boundary. The onset of the filamentation was, however, delayed. The likely cause is the high density gradient across the front, which alters the electron and ion skin depths and thus the characteristic scale of the filaments. The gradient is caused by the slowdown of the ions by the magnetic field of the EES and by the electron acceleration. The electrons are confined at the front in the direction of the shock normal so that they preserve the quasi-neutrality of the plasma, but they can move orthogonally to it. The latter results in a drift current. \\subsection{Field amplification} \\citet{MagAmp2} and \\citet{MagAmp1} have described observations of magnetic field amplification above the value expected from shock compression in SNRs. At the final simulation time the magnetic field energy density is increased in strength by over one order of magnitude, exceeding by far that expected from the magnetic field compression by the shock. A shock compresses only the magnetic field component perpendicular to the shock normal, which is weak in our case, and the amplification of its energy density can reach a factor of 4-7. The magnetic energy density at the simulation's end has been comparable to the box-averaged kinetic energy density in an interval spanning about 10-20 ion skin depths. The magnetic energy density due to the EES was at least twice as high as that of the background field in an interval covering 50 ion skin depths. Even if we take into account that the kinetic energy density close to the shock is increased by the accumulation of plasma, the magnetic energy density still constitutes a sizeable fraction of the local total plasma kinetic energy density. The EES has provided the main contribution to the magnetic energy density and exceeded that due to the filaments downstream by two orders of magnitude. Throughout this paper, we used normalized quantities in our simulation and we can scale the magnetic field amplitude to the relevant plasma conditions. If we set the electron density of the dense cloud to 1 cm$^{-3}$, we would obtain a peak magnetic field with a strength of 10 mG. However, we have to point out that our initial magnetic field amplitude has been higher than that expected for the ambient plasma, even if we take into account its amplification by cosmic ray-driven instabilities, and our simulation results may not be directly applicable. Where does the extra field come from? Amplification of the magnetic field can occur from the electron drift current arising from the $\\vec{E} \\times \\vec{B}$ drift motion in a layer close to the shock that is narrower than the ion gyroradius but wider than the electron gyroradius, see, e.g. \\citet{BaumjohannTreumann}. The current adds to the shock current and increases the jump in the perpendicular magnetic field. This can only occur when the ion and electron gyroradii differ, i.e. not in a pair plasma. The EES has a significant $E_x$-component and $|\\vec{B}_\\perp| \\approx |\\vec{B}_0|$. We thus obtain a $\\vec{E}\\times \\vec{B}$ drift orthogonal to the flow velocity vector. We have also found that the requirement to maintain quasi-neutrality of the plasma implies that the upstream electrons are dragged with the upstream ions across the EES, which moves with the shock. The resulting $\\vec{v}\\times\\vec{B}$ drift accelerates the electrons orthogonally to the shock propagation direction, further enhancing the net current and the magnetic field. Finally, magnetic fields of SLAMS are provided by the current due to the gyro-bunched ions, which rotate in the plane orthogonal to the wavevector. These mechanisms increase the mean magnetic field, and are different from the instability described by \\citet{MagAmp3} who has described a cosmic ray streaming instability which can amplify turbulent magnetic fields ahead of the shock. We can exclude Bell's instability here since we have not found energetic particles with a significant density moving upstream, which would provide the net current that drives this instability. \\subsection{Electron acceleration and upstream wave spectrum} The shock retains its planar structure after it forms. A circularly polarised large-scale precursor wave, the EES, expands into the foreshock. Its wavelength is several times the ion skin depth. It gradually rotates the quasi-parallel magnetic field into a quasi-perpendicular one at the current layer and it forces the incoming ions and electrons to interact with it nonlinearly. The ions are gyro-bunched and some of the incoming ions of the tenuous cloud are reflected by the forming shock. We have found evidence of a parametric instability \\citep{Instability} of the EES ahead of the foreshock and we could find at least two waves that may be the result of this parametric decay. These waves appear at late times, when the EES has expanded in space and is thus sufficiently monochromatic. They grow to an amplitude that introduces oscillations of the mean velocity of the ions of up to $c/5$. The interplay of the short-scale charge density waves and magnetowaves causes the electrons to be accelerated to highly relativistic speeds upstream of the forming shock. Similar electron acceleration (injection) mechanisms upstream of shocks involving whistler waves have been proposed by \\citet{Oblique3,Injection3,Gyrosurf}. The strongest electron acceleration is, however, observed at the location where the shock-reflected ion beam is developing. The electrons are accelerated to a peak Lorentz factor of 120 and their energy gain is thus comparable to the energy associated with the velocity change of the shock-reflected ions. If the electron acceleration is accomplished by the electromagnetic fields that reflect the incoming upstream ions, then the energy gain of the electrons may scale with the ion mass. We may expect in this case that the electrons are accelerated to a Lorentz factor $\\gamma_M \\approx 120 m_p / m_i$ that is $\\gamma_M \\approx 550$ if we would use the correct proton to electron mass ratio. It is interesting to see if this type of electron acceleration can also occur close to Solar system shocks. Let us consider the Earth bow shock as one of the best known collisionless shocks and let us assume a Solar wind speed of $4 \\times 10^5$ m/s to $7.5 \\times 10^5$ m/s. A specular reflection of the incoming Solar wind protons by the bow shock would change their energy by about 0.8 keV to 3 keV. Electrons with such energies are observed in a thin sheet close to the shock surface of perpendicular shocks \\citep{BowShock}. Perpendicular shocks are capable to produce shock-reflected ion beams \\citep{Shock6} and the electron acceleration mechanism we observe here may work also at the Earth bow shock. \\subsection{Future work} This simulation study was concerned with the collision of two plasma clouds at a mildly relativistic speed. Its purpose has been to better understand the conditions and the mechanisms involved in the formation of a shock. This shock will move at an essentially nonrelativistic speed below the initial collision speed and, thus, be relevant for fast SNR flows. A two-dimensional simulation geometry was necessary to assess the importance of the multi-dimensional filamentation instability for the shock dynamics. A quasi-parallel guiding magnetic field was used to slow down this filamentation, resulting in a planar (one-dimensional) shock. The formation of the shock could be observed, but the simulation had to be stopped at this time due to computational constraints. This simulation has, however, revealed several aspects that should be examined in more detail in more specialised simulation studies. The EES probably formed in response to our initial conditions. i.e. the sharp jump in the convection electric field at the cloud collision boundary. It has to be investigated if the EES also forms if this jump is decreased, for example by a higher-order field interpolation scheme or by a gradual decrease of the convection electric field by a smooth change in the plasma convection speed. The simulations by \\citet{Oblique1,Oblique2} suggest that this will leave unchanged the magnetic field amplification and the electron acceleration. However, the EES may not be so strong and coherent. The magnetic field amplitude in the present study is higher than it is realistic for a SNR scenario. Future studies must address how the shock formation depends on lower amplitudes of the guiding magnetic field. Computationally inexpensive parametric simulation studies that resolve only one spatial dimension may provide insight. Initial studies not discussed here indicate that the shock formation is delayed by a decreasing magnetic field amplitude. It is also necessary to follow the plasma collision for a longer time. An important aspect is here how far the EES can expand upstream and how strong the downstream magnetic field is. The planarity of the EES and of the shock boundary may permit us to use one-dimensional simulations, by which we can expand by at least an order of magnitude the box size along the collision direction. An one-dimensional simulation also allows us to examine with a larger number of particles per cell and, thus, lower noise levels the secondary instabilities driven by the EES. We have found evidence for an instability of the EES to a four-wave interaction. Lower noise levels would allow us to compare the amplitudes and phases of the EES with those of the secondary waves, which is necessary to demonstrate a coherent wave interaction. Extending the simulation time together with suitable initial conditions would, potentially, allow us to investigate what happens if the speed of the EES decreases below the local Alfv\\'en speed. It is possible that the EES decouples from the shock and propagates independently in form of an Alfv\\'en wave packet. Finally, it would be interesting to reduce the collision speed to about $c/10$, which is close to the expansion speed of the SNR shock, to see if and how many electrons are accelerated to relativistic speeds. This will provide insight into the electron injection efficiency of oblique shocks and, thus, into the ability of SNR shocks to accelerate electrons to cosmic ray energies." }, "0910/0910.5400_arXiv.txt": { "abstract": "{} {Recent studies have detected linear polarization in L dwarfs in the optical I band. Theoretical models have been developed to explain this polarization. These models predict higher polarization at shorter wavelengths. We discuss the polarization in the R and I band of 4 ultra cool dwarfs.} {We report linear polarization measurements of 4 ultra cool dwarfs in the R and I bands using the Intermediate dispersion Spectrograph and Imaging System (ISIS) mounted on the 4.2m William Herschel Telescope (WHT). } {As predicted by theoretical models, we find a higher degree of polarization in the R band when compared to polarization in the I band for 3/4 of these ultra cool dwarfs. This suggests that dust scattering asymmetry is caused by oblateness . We also show how these measurements fit the theoretical models. A case for variability of linear polarization is found, which suggests the presence of randomly distributed dust clouds. We also discuss one case for the presence of a cold debris disk. } {} ", "introduction": "A large number of ultra cool dwarfs have been detected in the last decade, and our understanding of these faint objects has kept improving. One of the challenging and fundamental aspects in the study of these objects is to understand the properties and distribution of condensate dust in the atmosphere. Observations of L dwarfs with effective temperatures of 1400-2200 K have led to the investigation of dust condensates in their atmospheres (\\cite {Kirkpatrick}; \\cite {Tsuji1996}). Because of complete gravitational settling, grains are expected to condense beyond the visible atmosphere for objects with effective temperatures below 1400 K (T-Dwarfs - \\cite{Allard 2001}; \\cite{Chabrier00}). At higher effective temperatures (1400-2200 K), grains can be present in the visible atmosphere because of incomplete gravitational settling (\\cite {Burrows and Sharp}; \\cite {Burrows 2001}; \\cite {Ackerman 2001}; \\cite {Allard 2001}; \\cite{Tsuji 2004}; \\cite{co03}; \\cite{helling} ). Recent discoveries of blue L dwarfs and L-T transition type dwarfs (as identified in \\cite {Knapp 2004}; \\cite {Chiu 2006}; \\cite{Tsuji03}) have brought forth models which could explain this phenomenon (eg. \\cite {BSH 2006}, \\cite {Knapp 2004}) by mechanisms which involve dust settling. It would be very important to validate these mechanisms. Linear polarization could be a very useful tool in understanding the observationally poorly constrained dust properties in the atmospheres of L dwarfs. The possibility of detecting polarization at optical wavelengths from grains in the atmospheres of L dwarfs was first raised by \\cite{SKR 2001}. Fast rotation of L dwarfs will induce the shape of their photosphere into the form of an oblate ellipsoid,(\\cite {Basri 2000}) and this nonsphericity will lead to the incomplete cancellation of the polarization from different areas of the stellar surface (\\cite {SKR 2001}). This prediction was first confirmed by the detection of linear polarization at 768 nm from a few L dwarfs by \\cite {Menard 2002}. Recently, \\cite{Zap 2005} have reported R and I band detection of linear polarization from several L dwarfs. Since polarization in the optical is unlikely to be due to Zeeman splitting of atomic or molecular lines or by synchrotron radiation, the observed polarization can be explained by single dust scattering in a rotationally induced oblate atmosphere (\\cite {S 2003}; \\cite{SK 2005}) or it could be due to large and randomly distributed dust clouds (\\cite{Menard 2002}). In this paper, we report polarization measurements of 3 L dwarfs (L0-L5) and one M9.5 dwarf with WHT/ISIS in both I and R bands . We also discuss our results comparing them with the recently published results of \\cite{goldman09}. Our measurements show the general trend that polarization is higher in the R band than the one in the I band. This trend strongly supports the presence of dust in the atmosphere of L dwarfs as it is very unlikely that any other mechanisms (such as the presence of magnetic field) can explain this observation at optical wavelengths (\\cite{Menard 2002}). We also discuss how the theoretical models (see Sect. 3) successfully fit our measured data. ", "conclusions": "\\begin{enumerate} \\item We report linear polarizaion measurements of 4 very nearby ultra cool dwarfs in the R and I bands. \\item We find that there is a trend (3 out of 4) of a higher degree of polarization at shorter wavelengths (R band) when compared to the I band as predicted by the theoretical models of \\cite{SK 2005}. \\item The L0 dwarf 2MASS J17312974+2721233 is interesting because of its relatively high polarization and requires follow-up studies. \\item We also fit theoretical models to predict the dust grain size and rotational velocities of three of the ultra cool dwarfs. \\item We find evidence for variability in the linear polarization for (2MASSW J1507476-162738). This suggests atmospheric activities like dynamical variations of the cloud cover in this object. \\end{enumerate}" }, "0910/0910.5546_arXiv.txt": { "abstract": "We report the detection of pulsations at 552~Hz in the rising phase of two type-I (thermonuclear) X-ray bursts observed from the accreting neutron star EXO~0748$-$676 in 2007 January and December, by the {\\it Rossi X-ray Timing Explorer}. % The fractional amplitude % was % 15\\% (rms). % The dynamic power density spectrum for each burst revealed an increase in frequency of $\\approx1$--2~Hz while the oscillation was present. The frequency drift, the high significance of the detections and the almost identical signal frequencies measured in two bursts separated by 11 months, confirms this signal as a burst oscillation similar to those found in 13 other sources to date. We thus conclude that the spin frequency in \\src\\ is within a few Hz of 552~Hz, rather than 45~Hz as was suggested from an earlier signal detection by \\cite{villarreal04}. Consequently, Doppler broadening must significantly affect spectral features arising from the neutron star surface, so that the narrow absorption features previously reported from an {\\it XMM-Newton}\\/ spectrum could not have arisen there. The origin of both the previously reported 45~Hz oscillation and the X-ray absorption lines is now uncertain. ", "introduction": "Neutron stars in low-mass X-ray binaries (LMXBs) provide observational evidence for their rapid spins reluctantly, and via increasingly diverse phenomena. Highly coherent burst oscillations, occuring only around the peak of thermonuclear (type-I) bursts, were the first such phenomenon to be discovered, and since have been detected in $\\approx14$ sources \\cite[e.g.][]{watts08}. Continuous pulsations in the persistent emission occur even more infrequently, and occur at just above the burst oscillation frequency in those sources which exhibit both \\cite[]{chak03a}. This result supports the hypothesis that the burst oscillation frequency traces the neutron star spin. Most recently, intermittent persistent pulsations have been detected in several sources, including one previously known burst oscillation source \\cite[e.g.][]{gal07a,altamirano07,casella07}. It is uncertain why some sources show pulsations or burst oscillations and others do not. Additionally, sources which exhibit burst oscillations do not do so in every burst. In fact, the presence of oscillations can be as rare as 1 burst in 14 \\cite[for 4U~1916$-$053; see][]{bcatalog}. The mechanism by which the burst oscillations are produced is yet another uncertainty. Although oscillations are observed at high (fractional) amplitudes early in some bursts, consistent with a spreading ``hot spot'' model \\cite[e.g.][]{stroh97b}, the oscillations in the burst tail, when the burning must have spread to the entire stellar surface, are harder to explain. The oscillations may instead (or also) arise from anisotropies in the surface brightness originating from hydrodynamic instabilities \\cite[]{slu02} or modes excited in the neutron star ocean (e.g. \\citealt{cb00}; see also \\citealt{heyl04,pb05}). The low-mass X-ray binary \\src\\ is particularly well-studied. This transient was discovered during {\\it EXOSAT}\\/ observations in 1985 \\cite[]{parmar86}, which also revealed the first thermonuclear bursts from the source as well as X-ray dipping activity. Synchronous X-ray and optical eclipses \\cite[]{crampton86} are observed once every 3.82~hr orbit. A variety of low- and high-frequency variability has been characterised \\cite[e.g.][]{homan99,homan00}, notably including a 695~Hz quasi-periodic oscillation. More recently, absorption features in the summed X-ray spectra of bursts observed by {\\it XMM-Newton}\\/ were identified as redshifted lines from near the neutron star surface \\cite[$z=0.35$;][]{cott02}. The narrowness of these features requires that the neutron star is rotating slowly, as rotation speeds $\\ga100$~Hz will broaden the line profiles to the point where they are undetectable \\cite[]{ozel03,chang06,bml06}. Subsequently, \\cite{villarreal04} detected a 45~Hz peak in the summed power spectrum of 38 thermonuclear bursts, which they interpreted as a spin frequency sufficiently slow to give negligible broadening. However, subsequent followup studies have failed to confirm the spectral line detection \\cite[e.g.][]{cott08}. Here we present analysis of recently observed bursts from \\src\\ which suggest that the neutron star spin is not 45~Hz, but 552~Hz. ", "conclusions": "\\label{sec2} There are substantial differences in the characteristics of the two burst oscillations (45~Hz and 552~Hz) now detected in \\src. The 45~Hz signal had a fractional amplitude of $\\approx3$\\% (rms) and was detected in the summed power spectrum of 38 bursts, rebinned by a factor of (typically) 64 or 128 to give a resolution of 1~Hz, calculated from light curves selecting photons within the energy range 6--60~keV, and covering intervals of typically 64 or 128~s of the burst decay. In contrast, the 552~Hz signal has a fractional amplitude of 15\\% (rms), was detected separately in unbinned power spectra calculated from 1- and 2-s light curves extracted over the full PCA energy range ($\\approx2$--60~keV) during the rise of two individual bursts, separated by 11~months. Additionally, the dynamic power density spectra give evidence for frequency evolution while the 552~Hz signal was present, similar to that observed in other burst oscillation sources. The key question is, which of these two signals traces the neutron star spin? We consider three alternatives. First, assuming both signals are genuine, the 45~Hz signal may arise from the neutron star spin, in which case the 552~Hz signal may arise from a high-order ($m\\approx12$--13) radial mode (\\citealt{cb00}; see also \\citealt{heyl04,pb05}). However, the appearance of this mode alone is difficult to explain; one would expect the excitation of many different modes, rather than just two with widely-separated orders. Second, if both signals are real but the 552~Hz signal instead indicates the neutron star spin, there is no known mechanism that can give rise to a 45~Hz signal from the neutron star surface. As suggested by \\cite{balman09}, the 45~Hz oscillation may instead arise in the boundary layer between the disk and neutron star. Third, the lack of a detection of the 45~Hz signal in the larger sample of bursts detected since 2004 suggests that the 45~Hz signal may have arisen from statistical fluctuations. Our analysis shows that including many more bursts in the power spectral sum has the effect only of reducing the detection significance to well below any conservative detection threshold. The signal power is also quite sensitive to other parameters of the data selection. Thus, while it is possible that the 45~Hz oscillation is transient and has not been detectable since, it is also difficult to rule out an origin in statistical noise with any confidence. The properties of the 552~Hz oscillation closely resemble those of the burst oscillations observed in other systems, that are believed to trace the neutron star spin to within a few Hz. In contrast, the properties of the 45~Hz oscillation are quite distinct, especially its low frequency and the inability to detect it in individual bursts. Thus, we conclude that the 552~Hz signal almost certainly traces the neutron star spin in \\src; since the signal appears to increase by up to 2~Hz during the burst rise, we expect that the true spin frequency may be a few Hz higher. The energy dependence of the 552~Hz oscillation amplitude in \\src\\ was similar to that of other burst oscillation sources \\cite[]{muno03c}, although the overall amplitudes were higher. The suggestion of the hard ($\\ga8$~keV) photons arriving earlier than the soft photons is however contrary to previous observations. In that respect we note that the previous analysis focussed principally on oscillations present throughout the tails of bursts, whereas the oscillations in \\src\\ were present only in the rise. It is not known whether the two types of oscillations have characteristically different variation of pulse arrival time with energy. The duty cycle of the 552~Hz oscillation (the number of bursts in which it was detected compared to the total number observed) is extremely small at $\\approx1.2$\\%, the smallest yet for all the burst oscillation sources (e.g. G08). The unprecedented scarcity of the oscillation raises the question of what was so unusual about the bursts that exhibited them. Compared to the entire sample of bursts observed by \\xte\\/ from \\src, the bursts on January 14 and December 13 had somewhat shorter timescales (calculated as the ratio of peak flux to fluence) $\\tau=12$--13~s, while the typical range is 15--30~s. Similarly, the rises were of shorter duration than average, at 2--4~s. Shorter burst timescales suggest a smaller proportion of hydrogen in the burst fuel than usual. Correspondingly, while the fluences were rather typical, the bursts reached higher than average peak fluxes. However, other bursts were observed with similar properties which did not exibit oscillations, so these properties do not uniquely determine their observeability. The January 14 burst occurred only 11.4~min after the previous event, and had a fluence only about 77\\% lower. Such short recurrence time bursts are common for \\src, and many examples have been detected by \\xte\\/ (e.g. G08) as well as {\\it XMM-Newton}\\/ \\cite[]{boirin07a}. However, the second burst in these pairs is typically much fainter than the first. Interestingly, it is also possible that the December 13 burst was the second in a closely-spaced pair, as a data gap (due to the 90~min satellite orbit) prevented observations until approximately 7~min before the event. Again, timing analysis of other such examples did not reveal any additional oscillations, however. We also compared the properties of the persistent emission at the time when the bursts with oscillations occurred, to other public \\xte\\/ observations of the source (from G08). The persistent flux in both observations was $\\approx3\\times10^{-10}\\ \\epcs$ (2.5--25~keV), approximately equal to the 50th percentile value of the flux distribution. Similarly, the hard and soft spectral colors\\footnote{Defined as in G08 as the ratio of the background-subtracted detector counts in the (8.6--18.0)/(5.0--8.6)~keV and the (3.6--5.0)/(2.2--3.6)~keV energy bands, respectively} were in the middle of the observed range, at $\\approx0.8$ and $\\approx1.6$, respectively. Thus, we found no evidence for an unusual spectral or intensity state at the time of the bursts with oscillations. A 552~Hz spin frequency for \\src\\ means that spectral features arising from the neutron star surface will be significantly Doppler broadened, as well as reducing the central line depth to only $\\la5$\\% of the continuum level \\cite[e.g.][]{cbw05}. Narrow lines from such rapidly-rotating objects can only arise if the system is viewed at extremely low inclinations; however, the eclipses and dipping activity in \\src\\ unambiguously indicate a high system inclination. Thus, the narrow ($\\approx0.1$~\\AA) spectral features reported by \\cite{cott02} could not arise from the neutron star surface. This conclusion is corroborated by other recent work, including the absence of lines in a deeper {\\it XMM-Newton}\\/ spectrum \\cite[]{cott08}, as well as difficulties explaining the inferred Fe column \\cite[]{cbw05}. However, the narrow spectral features are difficult to identify otherwise \\cite[e.g.][]{kong07}, and a satisfactory explanation remains elusive." }, "0910/0910.4540_arXiv.txt": { "abstract": "Following the recent discovery of $\\gamma$ rays from the radio-loud narrow-line Seyfert 1 galaxy PMN J0948+0022 ($z=0.5846$), we started a multiwavelength campaign from radio to $\\gamma$ rays, which was carried out between the end of March and the beginning of July 2009. The source displayed activity at all the observed wavelengths: a general decreasing trend from optical to $\\gamma-$ray frequencies was followed by an increase of radio emission after less than two months from the peak of the $\\gamma-$ray emission. The largest flux change, about a factor of about 4, occurred in the X-ray band. The smallest was at ultraviolet and near-infrared frequencies, where the rate of the detected photons dropped by a factor $1.6-1.9$. At optical wavelengths, where the sampling rate was the highest, it was possible to observe day-scale variability, with flux variations up to a factor of about 3. The behavior of PMN J0948+0022 observed in this campaign and the calculated power carried out by its jet in the form of protons, electrons, radiation and magnetic field are quite similar to that of blazars, specifically of flat-spectrum radio quasars. These results confirm the idea that radio-loud narrow-line Seyfert 1 galaxies host relativistic jets with power similar to that of average blazars. ", "introduction": "The recent detection by \\emph{Fermi Gamma-ray Space Telescope} of $\\gamma$ rays from the radio-loud narrow-line Seyfert 1 galaxy (RL-NLS1) PMN J0948+0022\\footnote{We note that the absolute magnitude of this source is $M_B=-23.6$, so formally matches also the definition of quasars.} ($z=0.5846$) opened new and interesting questions on the unified model of active galactic nuclei (AGN), the development of relativistic jets and the evolution of radio-loud AGN (Abdo et al. 2009a, Foschini et al. 2009a). Indeed, before \\emph{Fermi}/LAT (Large Area Telescope) it was known that $\\gamma$ rays from AGN are produced in blazars and radio galaxies, but we have to add also RL-NLS1s. NLS1s are active nuclei similar to Seyferts, where the optical permitted lines emitted from the broad-line region (BLR) are narrower than usual, with FWHM(H$\\beta$)$<2000$~km~s$^{-1}$ (see Pogge 2000, for a review). Other characteristics are [OIII]/H$\\beta<3$ and a bump of FeII, making them a peculiar class of AGN. NLS1s are different from Seyfert 2s, whose optical spectra typically display FWHM(H$\\beta$)$<1000$~km~s$^{-1}$, [OIII]/H$\\beta>3$ and no bump of FeII. NLS1s are also different from the naked AGN discovered by Hawkins (2004), a peculiar class of Seyferts without the BLR, which have [OIII]/H$\\beta>>3$. Indeed, NLS1s do have both BLR and the narrow-line region (NLR), but the BLR emits only permitted lines narrower than in Seyfert 1s (Rodr\\'iguez-Ardila et al. 2000). NLS1s are generally radio-quiet, but a small fraction of them ($<7$\\%, according to Komossa et al. 2006), are radio-loud. It is not clear how these sources fit into the framework of radio-loud AGN. Some studies of the average non-simultaneous multiwavelength properties (from radio to X-rays) of RL-NLS1s suggested some possibilities. Komossa et al. (2006) argued that RL-NLS1s could be some young stage of quasars, while Yuan et al. (2008) found some similarities to TeV BL Lacs, but having strong emission lines they would represent the so-called ``high-frequency peaked flat-spectrum radio quasars'' conjectured by Padovani (2007). Foschini et al. (2009b) found instead that there is no one-to-one correlation of RL-NLS1s properties with any specific type of blazar or radio galaxy. In some cases, there are similarities with flat-spectrum radio quasars, while others are like BL Lacs. Now, the first detection by \\emph{Fermi}/LAT of $\\gamma$ rays from one RL-NLS1 - namely PMN J0948+0022 - sets the definitive confirmation of the presence of a relativistic jet in these sources. The discovery of $\\gamma-$ray emission from other sources of this type (Abdo et al., in preparation) raise RL-NLS1s to the rank of $\\gamma-$ray emitting AGN. However, any average spectral energy distribution (SED) of a $\\gamma-$ray loud AGN leaves open several important questions on the mechanisms of radiation emission, such as whether the synchrotron self-Compton (SSC) or the external Compton (EC) production mechanism is dominant at high-energies and where the zone is where most of the dissipation occurs. Because PMN J0948+0022 is the first object of this new class of $\\gamma-$ray AGN, it is important to observe it for a long time, in order to understand if there is something unexpected and if its behavior is very different from blazars and radio galaxies or not. With these aims in mind, we decided to set up a multiwavelength campaign on this source. The campaign involved several space and ground-based facilities across the whole electromagnetic spectrum, from radio to $\\gamma$ rays (in alphabetical order): ATOM (Landessternwarte), F-GAMMA (Effelsberg), e-VLBI (EVN, LBA), \\emph{Fermi}, G. Haro Telescope (INAOE), Mets\\\"ahovi, OVRO, RATAN-600, \\emph{Swift}, SMARTS, MOJAVE (VLBA), WIRO. The period covered was between 2009 March 24 and July 5. We measured variability at multiple wavebands, modelled the resulting SEDs, and compared the results to those for more typical $\\gamma-$ray blazars in the FSRQ and BL Lac classes. Throughout this work, we adopted a $\\Lambda$CDM cosmology from the most recent \\emph{WMAP} results, which give the following values for the cosmological parameters: $h = 0.71$, $\\Omega_m = 0.27$, $\\Omega_\\Lambda = 0.73$ and with the Hubble-Lema\\^{i}tre constant $H_0=100h$ km s$^{-1}$ Mpc$^{-1}$ (Komatsu et al. 2009). ", "conclusions": "We thus confirm that PMN J0948+0022 -- despite being a radio-loud narrow-line Seyfert 1 -- hosts a relativistic $\\gamma-$ray emitting jet, similar to those of FSRQs, and confirms all the hypotheses adopted to model the non-simultaneous SED in Abdo et al. (2009a). This type of source can develop a relativistic jet like blazars and radio galaxies, even though the conditions of the environment close to its central spacetime singularity are quite different. This is indeed a new class of $\\gamma-$ray emitting AGN. We have shown that the variability at multiple wavebands and the physical parameters resulting from modelling the SEDs are typical of a source midway between FSRQs and BL Lacs. The calculated powers carried by the various components of the jet are low compared to the distributions of values for FSRQ, but above those of BL Lacs (cf Celotti \\& Ghisellini 2008, Ghisellini et al. 2009), and therefore within the average range of blazar powers, despite the relatively low mass of its black hole, $1.5\\times 10^{8}M_{\\odot}$ (Abdo et al. 2009a). The $\\gamma-$ray observations performed to date have not revealed very high fluxes, i.e. above the usual threshold adopted to define an outburst in normal blazars ($F_{E>100\\rm MeV}>10^{-6}$~ph~cm$^{-2}$~s$^{-1}$). However, it is not clear yet if this is due to the duty cycle of this source -- and hence if we have just observed a minor event -- or if the different environmental conditions in the core of RL-NLS1s hampers the development of a high power jet. This question will likely be answered by the continuous monitoring that \\emph{Fermi}/LAT is performing on this and other sources of this type." }, "0910/0910.4892_arXiv.txt": { "abstract": "{} { We used VLT/VIMOS images in the $V$ band to obtain light curves of extrasolar planetary transits OGLE-TR-111 and OGLE-TR-113, and candidate planetary transits: OGLE-TR-82, OGLE-TR-86, OGLE-TR-91, OGLE-TR-106, OGLE-TR-109, OGLE-TR-110, OGLE-TR-159, OGLE-TR-167, OGLE-TR-170, OGLE-TR-171. } { Using difference imaging photometry, we were able to achieve millimagnitude errors in the individual data points. We present the analysis of the data and the light curves, by measuring transit amplitudes and ephemerides, and by calculating geometrical parameters for some of the systems\\thanks{Photometry of the transiting objects is available at the CDS via anonymous ftp to cdsarc.u-strasbg.fr}. } { We observed 9~OGLE objects at the predicted transit moments. Two other transits were shifted in time by a few hours. For another seven objects we expected to observe transits during the VIMOS run, but they were not detected. } { The stars OGLE-TR-111 and OGLE-TR-113 are probably the only OGLE objects in the observed sample to host planets, with the other objects being very likely eclipsing binaries or multiple systems. In this paper we also report on four new transiting candidates which we have found in the data. } ", "introduction": "The field of extrasolar planets is developing rapidly, producing exciting results at an accelerated pace. The discovery of the first extrasolar ``hot Jupiter'' around the nearby solar-type star 51~Peg using precise radial velocity measurements \\citep{may95} spurred a number of discoveries. Chief among these was the discovery of transits around the nearby solar-type star HD~209458 \\citep{char00,hen00}. The success of the radial velocity studies also boosted extrasolar planetary searches using other techniques such as microlensing and transits. Currently more than 60 transiting extrasolar planets are known\\footnote{See http://exoplanets.eu/}. Many candidates were discovered by the OGLE team who carried out systematic searches, monitoring millions of stars along fields located in the Milky Way disk \\citep{uda02a,uda02b,uda03,pont08}. Of more than 200 transiting candidates, already seven OGLE transits have been confirmed as being due to planets: OGLE-TR-10 \\citep{kon05}, OGLE-TR-56 \\citep{kon03}, OGLE-TR-111 \\citep{pont04}, OGLE-TR-113 \\citep{bou04,kon04}, OGLE-TR-132 \\citep{bou04}, OGLE-TR-182 \\citep{pont08}, OGLE-TR-211 \\citep{uda08}. Most of the other targets are eclipsing low-mass stars or brown dwarfs, or due to blends of normal stars, triplets, etc. Why such interest in observing transiting candidates? The radial velocities give orbital parameters such as period, semi-major axis, and projected mass ($M~{\\rm sin}~i$). The transits give not only the orbital parameters like period and inclination, but also the planet sizes: the eclipse amplitude is simply $(r/R)^2$. Thus, from combined radial velocities and transits we know the mass of the planet without the inclination ambiguity, and the radius, which gives a density. The few planets so studied appear to be indeed inflated gaseous planets, i.e. ``hot Jupiters''. One difficulty is that the derived planet radii are only good to 10-15\\%. More accurate transit photometry is needed to improve those estimates, as argued by \\cite{mou04}. For example the discoveries by \\cite{pont05a,pont05b} of planet-sized stars around OGLE-TR-106 and OGLE-TR-122, help to constrain the models of planetary systems. These planets are under intense irradiation, which inflates their sizes, depending on their own orbital parameters and intrinsic characteristics \\citep{bar03,burr02}. In some way the OGLE planets constitute the extreme cases, because of their very short periods. We conducted a programme to monitor photometrically the OGLE transit candidates, and here we present precise photometry for these transits. Some of the objects, namely OGLE-TR-109, OGLE-TR-111, OGLE-TR-113 have been already analyzed by \\cite{fer06}, \\cite{min07}, \\cite{diaz07}, respectively. In this paper we present an analysis of all OGLE transits in the VIMOS images. We also have searched for new transits. Section~2 gives details on the observations, selection and properties of the sample. Section~3 describes reductions of the data. In sections~4-8 we present the results obtained from the observed light curves of the OGLE transits. Section~9 describes new transiting candidates we have found in the data. Finally, section~10 states our main conclusions. ", "conclusions": "$V$-band images from VLT/VIMOS were used to obtain light curves of extrasolar planetary transits: OGLE-TR-111, OGLE-TR-113, and candidate planetary transits: OGLE-TR-109, OGLE-TR-159, OGLE-TR-167, OGLE-TR-170, OGLE-TR-171. With difference imaging photometry we were able to achieve millimagnitude errors in the individual data points. The following seven OGLE transits were recorded as full events: 106, 109, 111, 113, 159, 170, 171. Four transits were detected as partial events: 86, 91, 110, 167. All full and partial transits but OGLE-TR-91 and OGLE-TR-109 were observed at the predicted transit times. No transits were recorded for 19 objects. Based on the shape of the obtained light curves and some results from spectroscopic follow-up studies we show that the objects OGLE-TR-111 and OGLE-TR-113 are probably the only OGLE stars in the sample to host planets. In the paper we also report on four new transiting candidates we have found in the VIMOS data. One of the events, transit-4, with the duration time of about 3.7~h and the amplitude of about 0.02~mag, is a particularly good candidate for a planetary transit. Faintness of the object (17.8~mag in $V$) may severely hamper spectroscopic verification. However, all four candidates require photometric follow-up studies to look for their periodic nature first." }, "0910/0910.3683_arXiv.txt": { "abstract": "The next generation of proposed galaxy surveys will increase the number of galaxies with photometric redshift identifications by two orders of magnitude, drastically expanding both the redshift range and detection threshold from the current state of the art. Obtaining spectra for a fair sub-sample of this new data could be cumbersome and expensive. However, adequate calibration of the true redshift distribution of galaxies is vital to tapping the potential of these surveys to illuminate the processes of galaxy evolution, and to constrain the underlying cosmology and growth of structure. We examine here a promising alternative to direct spectroscopic follow up: calibration of the redshift distribution of photometric galaxies via cross-correlation with an overlapping spectroscopic survey whose members trace the same density field. We review the theory, develop a pipeline to implement the method, apply it to mock data from N-body simulations, and examine the properties of this redshift distribution estimator. We demonstrate that the method is generally effective, but the estimator is weakened by two main factors. One is that the correlation function of the spectroscopic sample must be measured in many bins along the line of sight, which renders the measurement noisy and interferes with high quality reconstruction of the photometric redshift distribution. Also, the method is not able to disentangle the photometric redshift distribution from evolution in the bias of the photometric sample. We establish the impact of these factors using our mock catalogs. We conclude it may still be necessary to spectroscopically follow up a fair subsample of the photometric survey data. Nonetheless, it is significant that the method has been successfully implemented on mock data, and with further refinement it may appreciably decrease the number of spectra that will be needed to calibrate future surveys. ", "introduction": "\\label{sec:intro} It is essential to calibrate the true redshift distribution of galaxies in a photometric survey if the survey is to be utilized to its full potential. One application of survey data that requires a detailed understanding of the distribution of galaxies is weak gravitational lensing tomography. The shearing of the shapes of distant galaxies via weak gravitational lensing is a powerful cosmological probe that can be used to study the distribution of dark matter, the nature of dark energy, the formation of large scale structures in the universe, as well as fundamental properties of elementary particles and potential modifications to the general theory of relativity (recent studies include \\cite{2008ARNPS..58...99H}, \\cite{2009MNRAS.395..197T}, \\cite{2009A&A...497..677K}, \\cite{2009PhRvD..79b3520I}). Cosmic shear measurements statistically examine minute distortions in the orientations of high redshift galaxies, whose shapes have been sheared by intervening dark matter structures. Although weak gravitational lensing provides only an integrated measure of the intervening density field, using source populations at different redshifts permits some degree of three dimensional reconstruction, known as tomography. The distortions are small (at the 1\\% level) and the intrinsic orientation of the source galaxies is unknown, thus large galaxy samples are required to map the density field and probe the growth of density fluctuations with precision. Existing cosmic shear measurements have already constrained the amplitude of the dark matter fluctuations at the 10\\% level \\cite{2004MNRAS.347..895H} \\cite{2004ApJ...605...29R} \\cite{2005MNRAS.359.1277M} \\cite{2006A&A...452...51S} and there are many exciting galaxy survey proposals that will increase the available number of source galaxies by two orders of magnitude including DES, DUNE, Euclid, LSST, PanStarrs, SNAP, and Vista. Because these large galaxy surveys will have photometric rather than spectroscopic redshift identifications, the community has carefully attended to fine tuning the calibration of the photometric redshifts, minimizing biases and catastrophic errors \\cite{2008MNRAS.386..781M}, \\cite{2008MNRAS.390..118L}, \\cite{2009MNRAS.398.2012F}, \\cite{2009AJ....138...95X}, \\cite{2009arXiv0908.4085G}. Unlike experiments that use the galaxy positions to directly trace the underlying dark matter distribution, such as baryon acoustic oscillation studies, weak lensing analyses do not require a precise redshift identification for each individual source galaxy. It is sufficient to accurately determine the redshift distribution of the sources. However, lensing measurements are extremely sensitive both to error and bias in the source distribution \\cite{2002PhRvD..65f3001H}. Attaining an accurate source distribution will be crucial if weak lensing measurements are to be competitive with other cosmological probes in constraining the cosmological parameters. Another example where calibration of the true distribution of galaxies may be essential is in using the abundances and clustering of different galaxy populations to connect galaxies at late times to their potential progenitors at early times (as in e.g. \\cite{2009ApJ...696..620C}). Such studies also utilize the luminosity functions of galaxies in each redshift slice, and sometimes divide these into different rest-frame color bins. To avoid potential systematics in inferences made about galaxy evolution, it will be necessary to know if some fraction of the population in a given photometric redshift bin is actually living at a different redshift, especially if there is an asymmetry in such errors that depends on color. Recently, an alternative approach to attaining an accurate source redshift distribution has been proposed in \\cite{2008ApJ...684...88N}. This method is similar to cross-correlation techniques used in \\cite{1985MNRAS.212..657P} and \\cite{2006ApJ...644...54M}, and the idea has also been studied theoretically in \\cite{2006ApJ...651...14S} and \\cite{2009arXiv0902.2782B}. A similar technique was used in \\cite{2009A&A...493.1197E} to check the redshift distribution for interlopers. Similar in spirit, the analysis of \\cite{2009arXiv0910.2704Q} uses close angular pairs of photometric galaxies to constrain the photometric errors without the use of a spectroscopic sample. The cross-correlation method determines the photometric redshift distribution by utililizing the cross correlation of the galaxies in the photometric sample with an overlapping spectroscopic sample that traces the same underlying density field. One advantage of this approach is that the spectroscopic sample used to calibrate the photometric redshift distribution can be comprised of bright rare objects such as quasars or Luminous Red Galaxies (LRGs) whose spectra are relatively easy to obtain, and indeed may already exist in legacy data. Spectra could also be obtained for emission line galaxies (ELGs), which are easy to follow up but may not represent a fair subsample. Another advantage is that catastrophic redshift errors in the photometry do not systematically bias the redshift distribution estimate, they merely contribute to the noise. The cross-correlation method makes use of two observables, the line-of-sight projected angular cross-correlation between the photometric and spectroscopic samples $w_{ps}(\\theta)$, and the three dimensional autocorrelation function of the spectroscopic sample $\\xi_{ss}(r)$. By postulating a simple proportionality between the autocorrelation function of the spectroscopic objects and the three dimensional cross-corelation function between the two samples $\\xi_{ps}(r) \\propto \\xi_{ss}(r)$, it is potentially possible to infer a very accurate redshift distribution for the photometric sample. This assumption is guaranteed to be valid if the spectroscopic sample is a sub-sample of the photometric population, but may be problematic if the two sets of tracers have different bias functions with respect to the dark matter. In this paper we develop a pipeline to apply the cross-correlation method. In section \\ref{sec:theory} we review the theory and explain how the method works. We highlight its strengths and examine potential drawbacks and systematic effects. As a proof of concept, in section \\ref{sec:sims} we use the halo model to populate N-body simulations with mock photometric and spectroscopic galaxy data to quantify the properties of this redshift distribution estimator. We examine the extent to which different bias functions interfere with the reconstruction of the true distribution of photometric galaxies. In section \\ref{sec:disc} we discuss inherent tradeoffs, outline outstanding theoretical questions, and draw our conclusions. We leave a more detailed discussion of error propagation to the appendix. ", "conclusions": "\\label{sec:disc} We have outlined the theory behind the cross-correlation method for calibrating the redshift distribution of objects with photometric redshifts, and developed a pipeline than can be used to apply the method to survey data. We have created mock simulations to test the pipeline. We have succeeded in reconstructing the redshift distribution of the mock photometric galaxies using the angular cross correlation of these galaxies with an overlapping spectroscopic sample (whose redshifts are known). We have not used any redshift information about the photometric sample. We have demonstrated the validity of the method. We have also identified the aspects that are likely to be the limiting factors in relying upon this method to provide accurate redshift distribution information. These limiting factors are 1) that the spectroscopic sample must be binned along the line of sight, causing their correlation functions to be noisy and interfering with the reconstruction, and 2) that the bias evolution of the photometric sample cannot be disentangled from the redshift distribution that is reconstructed, which may force the necessity for the follow up of a fair subsample of galaxies. Improved modeling of theoretical priors could yield large dividends if these factors can be mitigated. The analysis has revealed a number of trade-offs that exist in application of this method. For a given set of calibrators, there is a trade-off between the resolution (bin spacing) of the redshift distribution reconstruction, and the error bars on any individual point. Thus if a population of catastrophic outlier were discovered, for example, it would be interesting to investigate whether it is more important to know how many there are, or at exactly which redshift they lie. We also find a trade-off between the number of calibrators and the quality of the reconstruction. As the number of available spectra are decreased, the spectroscopic redshift bins need to widen to maintain equivalent signal strength. This in turn affects how finely the reconstructed redshift distribution can be sampled. Bimodal behavior may be lost, and with insufficient priors (such as the smoothness prior we implemented here), false bimodal behavior may appear in the form of anti-correlated adjacent points. Widening the bins to compensate for fewer spectra also introduces another difficulty. Recall that the error in the angular cross-correlation function goes as $\\delta w_{ps} \\sim 1/ \\sqrt{n_{\\rm pairs}}$. This means that \\begin{eqnarray} \\frac{\\delta w_{ps}}{w_{ps}}=\\frac{1}{\\sqrt{n_{\\rm pairs}}} \\left[ \\int_0^\\infty \\xi_{ps} \\phi_p d\\chi\\right]^{-1} \\end{eqnarray} For the purposes of this order of magnitude argument, suppose $\\phi_{p}$ were constant, then to integrate to 1 requires $\\phi_{p}(\\chi) \\propto 1/\\Delta\\chi_{\\rm rb}$. The subscript rb is used to indicate the width of the reconstruction bin (not the spectroscopic bins). When we remove it from the integral we are left with an expression that is $w_p(r_p)$ \\begin{eqnarray} \\frac{\\delta w_{ps}}{w_{ps}}=\\frac{1}{\\phi_p \\sqrt{n_{\\rm pairs}}} \\left[ \\int_0^\\infty \\xi_{ps} d\\chi\\right]^{-1} \\nonumber \\\\ \\sim \\frac{\\Delta\\chi_{\\rm rb}}{\\phi_p w_p(r_p) \\sqrt{n_{\\rm pairs}}} \\end{eqnarray} If the reconstruction bin is widened by a factor of two, the number density of photometric galaxies will have to be increased by a factor of 4 to preserve the same accuracy in the cross correlation measurement. Therefore in this method there is also a trade-off between the number of spectra that the survey can afford versus the number of photometric galaxies they can afford. In this analysis we have not made any use of the photometric redshifts, which although not sufficiently accurate to determine the true redshift distribution of the population, still contain significant information about it. Modern techniques such as in \\cite{2009arXiv0908.4085G} have made it possible to assign a probability distribution for the redshift of each individual galaxy in a photometric survey, rather than a single best estimate and error bar. We propose that combining these probabilities for all the galaxies in a given redshift bin constitutes a reasonably powerful prior, that can take the place of the smoothing we have introduced in this analysis. This is fortunate because the solution is frequently wrecked by the smoothing criterion, although the analysis cannot be performed without it. Another detail that we leave for a future study is that the spectroscopic sample need not be comprised of a single rare population, it is quite conceivable to target certain regions of redshift space that are known to be problematic more heavily. It must also be possible to use less rare tracers in regions with smaller volume. We leave such optimization to the future. There are a few important factors that we have neglected in this analysis. As pointed out by \\cite{2009arXiv0902.2782B}, weak gravitational lensing will induce correlations between the positions of calibrator galaxies in the foreground with photometric galaxies in the background, and vice versa. This will need to be carefully controlled for the method to reliably calibrate redshift distributions. We also have not mentioned the complication of the integral constraint, which as discussed in e.g. \\cite{2007APh....26..351H} can lead to very significant errors when the volume and the scales in the correlation function are of comparable size. This may be as significant an issue as evolution in the large scale bias of the photometric population, though it may be mitigated if integral constraint errors are correlated between photometric and spectroscopic samples. Fortunately, improved estimators exist (in \\cite{2007MNRAS.376.1702P} for example) and should certainly be incorporated into the pipeline. Having now demonstrated that the method is viable, with further refinement the cross-correlation method applied in conjunction with direct follow up surveys may significantly reduce the number of spectra that are required to calibrate photometric redshift distributions to the desired accuracy. There are a few other advantages that we have not touched upon in detail. As we have shown, the cross correlation method can be used to calibrate the redshift distribution all the way along the line of sight, and as such is uniquely suited to detection and calibration of catastrophic redshift errors, even if these errors are so rare that they are missed by conventional follow up. Also, the reconstruction should not be adversely affected by redshift deserts, regions where no spectroscopic redshifts are available. We are optimistic that this technique will provide a useful complementary approach to conventional calibration techniques, and every effort should be made to refine the method further." }, "0910/0910.3010_arXiv.txt": { "abstract": "Measurement of the spatial distribution of neutral hydrogen via the redshifted 21~cm line promises to revolutionize our knowledge of the epoch of reionization and the first galaxies, and may provide a powerful new tool for observational cosmology from redshifts $1d_\\mathrm{wd}$ for $\\sim$1/5 of the systems, a tendency found already in our previous work. The hypothesis that magnetic activity raises the temperature of the inter-spot regions in active stars that are heavily covered by cool spots, leading to a bluer optical colour compared to inactive stars, remains the best explanation for this behaviour. We also make use of SDSS-GALEX-UKIDSS magnitudes to investigate the distribution of WDMS binaries, as well as their white dwarf effective temperatures and companion star spectral types, in ultraviolet to infrared colour space. We show that WDMS binaries can be very efficiently separated from single main sequence stars and white dwarfs when using a combined ultraviolet, optical, and infrared colour selection. Finally, we also provide radial velocities for 1068 systems measured from the \\Lines{Na}{I}{8183.27,8194.81} absorption doublet and/or the H$\\alpha$ emission line. Among the systems with multiple SDSS spectroscopy, we find five new systems exhibiting significant radial velocity variations, identifying them as post-common-envelope binary candidates. ", "introduction": "\\label{s-intro} Binaries containing a white dwarf primary plus a main sequence companion were initially main sequence binaries in which the more massive star evolved through the giant phase and became a white dwarf. In the majority of cases the initial separation of the main sequence binary is wide enough to allow the evolution of both stars as if they were single. A small fraction is believed to undergo a phase of dynamically unstable mass transfer once the more massive star is on the giant branch or the asymptotic giant branch \\citep{webbink84-1, dekool92-1, willems+kolb04-1}. As a consequence of this mass-transfer the envelope of the giant will engulf its core and the companion star, i.e. the system is entering a common envelope phase \\citep[CE, e.g.][]{livio+soker88-1, iben+livio93-1, taam+sandquist00-1, webbink07-1}. Friction inside this envelope causes a rapid decrease of the binary separation. Henceforth orbital energy and angular momentum are extracted from the binary orbit and lead to the ejection of the envelope, exposing a post-common-envelope binary (PCEB). As a consequence the orbital period distribution of WDMS binaries is clearly bi-modal, with PCEBs concentrated at short orbital periods, and wide WDMS binaries (non-PCEBs) at long orbital periods \\citep{willems+kolb04-1}. After the ejection of the envelope, close WDMS binaries evolve to shorter orbital periods through angular momentum loss via magnetic braking and/or through the emission of gravitational waves. WDMS binaries include progenitors of a wide range of astronomical objects such as cataclysmic variables and super-soft X-ray sources, with some of those objects likely evolving at later stages into type Ia supernova. Population synthesis models have been developed for a variety of binary stars undergoing CE evolution \\citep{dewi+tauris00-1, nelemans+tout05-1, politano+weiler06-1, politano+weiler07-1, davisetal08-1, davisetal09-1}. However, the theoretical understanding of both CE evolution and magnetic braking is currently poorly constrained by observations \\citep{schreiber+gaensicke03-1}, and progress on this front is most likely to arise from the analysis of a large sample of PCEBs that are well-understood in terms of their stellar components. WDMS binaries appear most promising in that respect, as their stellar components are relatively simple, and the SDSS \\citep{yorketal00-1, stoughtonetal02-1, abazajianetal09-1} offers the possibility to dramatically increase the number of WDMS binaries available for detailed follow-up studies. Up to date there exist four compilations of SDSS WDMS binaries, namely \\citet{eisensteinetal06-1} (obtained as part of their search of white dwarfs), \\cite{silvestrietal07-1} (which contains also \\citet{raymondetal03-1,silvestrietal06-1} as a subset, and was claimed to be complete for the SDSS DR5), \\citet{augusteijnetal08-1} (obtained from SDSS DR5 using colour cuts plus proper motions), and \\citet{helleretal09-1} (obtained from a search for sdB stars in the SDSS DR6). Here we initiate a comprehensive study to compile a master sample of spectroscopic WDMS binaries from SDSS. In this first publication we make use of the SDSS spectroscopic DR6 \\citep{adelman-mccarthyetal08-1} to create a catalogue of 1602 WDMS binaries and candidates that were serendipitously observed. This list is not only a superset of all those previous SDSS WDMS binary compilations, but also includes 440 new WDMS binaries. In addition we provide a coherent analysis of the system parameters of both stellar components, as well as an extension to GALEX/UKIDSS colours to study the properties of SDSS WDMS binaries. In a parallel effort, and as part of SEGUE (the SDSS Extension for Galactic Understanding and Exploration), some of us carried out a dedicated program to identify $\\sim300$ WDMS binaries containing cold white dwarfs \\citep{schreiberetal07-1}, a population clearly underrepresented in previous samples of WDMS binaries. A detailed analysis of the SEGUE population of WDMS is in preparation (Schreiber et al. 2009). Finally, we plan to summarise our efforts by presenting the WDMS binary content of the final SDSS data release \\citep[DR7,][]{abazajianetal09-1}. This last paper will be accompanied by a public online data base of WDMS binaries. The final master sample of ``all'' spectroscopic SDSS WDMS binaries will form a superb database for future follow-up studies of WDMS binaries. In particular, analysing the fraction and the characteristics of the PCEBs among the WDMS binaries may provide strong constraints on current theories of compact binary evolution \\citep{rebassa-mansergasetal07-1, schreiberetal08-1, rebassa-mansergasetal08-1, pyrzasetal09-1, nebotetal09-1, schwopeetal09-1}. The structure of the paper is as follows. In Sect.\\,\\ref{s-ident} we present our method of identifying WDMS binaries, and estimate the completeness of the sample in Sect.\\,\\ref{s-completeness}. In Sect.\\,\\ref{s-final} we provide our final catalogue. Using a spectral decomposition/model atmosphere analysis, we derive white dwarf effective temperatures, surface gravities, masses, companion star spectral types, and distances in Sect.\\,\\ref{s-param}. We compare these stellar parameters to those obtained in other studies in Sect.\\,\\ref{s-comparison}, and study the selection effects of WDMS binaries within SDSS in Sect.\\,\\ref{s-effects}. We provide then colour-colour diagrams and colour-colour cuts for WDMS binaries in Sect.\\,\\ref{s-colors} and Sect.\\,\\ref{s-cuts}, respectively. We finally measure the \\Lines{Na}{I}{8183.27,8194.81} absorption doublet and/or the H$\\alpha$ emission radial velocities in Sec.\\,\\ref{s-rvs}, and summarise our results in Sect.\\,\\ref{s-conc}. \\begin{figure*} \\begin{center} \\includegraphics[angle=-90,width=\\textwidth]{wdms.ps} \\caption{Six examples of previously known WDMS binaries used in this work as WDMS binary templates. SDSS names and MJD-PLT-FIB identifiers are indicated for each of them. White dwarf effective temperatures and spectral type of the companions (see Sec.\\,\\ref{s-param}) are also indicated in each panel.} \\label{f-templ} \\end{center} \\end{figure*} ", "conclusions": "\\label{s-conc} We have presented a catalogue of 1602 WDMS binaries from the spectroscopic SDSS DR6. We have used a decomposition/fitting technique to measure the effective temperatures, surface gravities, masses and distances to the white dwarfs, as well as the spectral types and distances to the companions in our catalogue. Distributions and density maps obtained from these stellar parameters have been used to study both the general properties and the selection effects of WDMS binaries in SDSS. A comparison between the distances measured to the white dwarfs and the main sequence companions shows $d_\\mathrm{sec}>d_\\mathrm{wd}$ for $\\sim$1/5 of the systems. We also have made use of GALEX, SDSS and UKIDSS magnitudes to study the distribution of WDMS binaries in colour-colour space and present simple colour-cuts that allow to clearly separate WDMS binaries from other stellar objects. Finally, we have measured radial velocities for 1068 WDMS binaries measured from the \\Lines{Na}{I}{8183.27,8194.81} absorption doublet and/or the H$\\alpha$ emission line. Among the systems with multiple SDSS spectroscopy, we find five new WDMS binaries showing significant radial velocity variations identifying them as PCEB candidates. The here presented new, updated, and most complete catalogue of WDMS binaries from the SDSS represents a superb data base for future follow-up studies that may significantly contribute to a better understanding of close compact binary star evolution." }, "0910/0910.1223_arXiv.txt": { "abstract": "We report the results of a search for 12.2-GHz methanol maser emission, targeted towards 113 known 6.7-GHz methanol masers associated with 1.2-mm dust continuum emission. Observations were carried out with the Australia Telescope National Facility (ATNF) Parkes 64-m radio telescope in the period 2008 June 20 - 25. We detect 68 12.2-GHz methanol masers with flux densities in excess of our 5-$\\sigma$ detection limit of 0.55~Jy, 30 of which are new discoveries. This equates to a detection rate of 60 per cent, similar to previous searches of comparable sensitivity. We have made a statistical investigation of the properties of the 1.2-mm dust clumps with and without associated 6.7-GHz methanol maser and find that 6.7-GHz methanol masers are associated with 1.2-mm dust clumps with high flux densities, masses and radii. We additionally find that 6.7-GHz methanol masers with higher peak luminosities are associated with less dense 1.2-mm dust clumps than those 6.7-GHz methanol masers with lower luminosities. We suggest that this indicates that more luminous 6.7-GHz methanol masers are generally associated with a later evolutionary phase of massive star formation than less luminous 6.7-GHz methanol maser sources. Analysis of the 6.7-GHz associated 1.2-mm dust clumps with and without associated 12.2-GHz methanol maser emission shows that clumps associated with both class~II methanol maser transitions are less dense than those with no associated 12.2-GHz methanol maser emission. Furthermore, 12.2-GHz methanol masers are preferentially detected towards 6.7-GHz methanol masers with associated OH masers, suggesting that 12.2-GHz methanol masers are associated with a later evolutionary phase of massive star formation. We have compared the colours of the GLIMPSE point sources associated with the maser sources in the following two subgroups: 6.7-GHz methanol masers with and without associated 12.2-GHz methanol masers; and 6.7-GHz methanol masers with high and and those with low peak luminosities. There is little difference in the nature of the associated GLIMPSE point sources in any of these subgroups, and we propose that the masers themselves are probably much more sensitive than mid-infrared data to evolutionary changes in the massive star formation regions that they are associated with. We present an evolutionary sequence for masers in high-mass star formation regions, placing quantitative estimates on the relative lifetimes for the first time. ", "introduction": "Interstellar masers are one of the most readily observed signposts of star formation, particularly emission from the hydroxyl (OH), water (H$_2$O) and methanol (CH$_3$OH) molecules as they are relatively common and intense. Because these masers arise at radio frequencies, they are not affected by the dense optical-obscuring gas and dust present at the early stages of massive star formation thereby allowing us to probe these stars at the earliest stages. The masers are powerful probes of the kinematics of the star formation regions, but we are only recently beginning to be able to use the masers to investigate aspects such as the physical conditions through comparison of observational data and maser pumping theories \\citep{Cragg01,Sutton01}. The different maser species favour different physical conditions and are thought to trace different evolutionary phases of massive star formation \\citep{Ellingsen07}. Since many sources show emission from multiple maser transitions or species there must be significant overlap for the evolutionary phase traced by the most common types of masers. Interstellar masers have been detected towards more than a thousand star formation regions in our Galaxy from transitions of methanol, hydroxyl and water. Transitions of class II methanol masers have some advantages over water and OH masers as they exclusively trace sites of massive star formation \\citep{Minier03,Xu08} while water and OH are known to also be associated with other astrophysical objects such as evolved stars and supernova remnants. Two of the strongest maser species found towards star formation regions are transitions of methanol at 6.7- and 12.2-GHz, both of which are known to trace an early evolutionary stage of massive star formation \\citep{Ellingsen06}. Complementary observations of these two strong methanol maser transitions are especially useful in probing the physical conditions of the environments in which they arise, as they are known to typically be co-spatial \\citep{Menten92,Norris93}. To date there have been several searches for 12.2-GHz methanol maser emission towards 6.7-GHz methanol maser emission \\citep*[e.g.][]{Caswell95b,Blas04}. \\citet{Caswell95b} targeted their search for 12.2-GHz methanol masers towards 238 6.7-GHz methanol masers that had been detected towards sites of OH masers and achieved a detection rate of 55 per cent. \\citet{Blas04} directed their observations at 12.2-GHz towards 6.7-GHz methanol masers that were observed towards a sample of {\\em IRAS} (Infrared Astronomy Satellite) selected sources \\citep{Szy00} as well as 6.7~GHz methanol masers that were detected in a blind search that had relatively poor positional accuracy and sensitivity \\citep{Szy02}. Searches such as those by \\citet{Caswell95b} and \\citet{Blas04} have shown that 12.2-GHz methanol maser emission is only rarely brighter than the associated 6.7-GHz methanol maser emission. This coupled with the fact that there have been no serendipitous detections of 12.2-GHz masers without a 6.7-GHz counterpart mean that it is unlikely that an unbiased Galactic survey would uncover many, if any, more 12.2-GHz methanol masers than would be yielded from a search targeted towards 6.7-GHz methanol maser sources. Recently, a small subset of the strong 6.7-GHz methanol masers from the sources observed by \\citet{Hill05} at 1.2-mm, were searched for the presence of 12.2-GHz methanol masers with the University of Tasmania 26~m Mount Pleasant radio telescope \\citep{Lewis07}. \\citet{Lewis07} detected 17 12.2-GHz methanol masers towards the 27 sites searched with a detection limit of $\\sim$4~Jy. \\citet{Hill05} observed 131 regions that were suspected of undergoing massive star formation, using the presence of previously identified 6.7-GHz methanol masers and/or an ultracompact \\ionhy region to select these regions. A total of 404 1.2-mm dust clumps were identified, with 113 of these associated with 6.7-GHz methanol masers, 35 of which are associated with radio continuum emission \\citep{Walsh98}. Statistical analysis by \\citet{Lewis07} of the properties of 1.2-mm dust clumps associated with 6.7-GHz methanol maser emission with and without 12.2-GHz methanol masers showed that those 1.2-mm dust clumps with both 6.7-GHz and 12.2-GHz were less massive and less dense than the 1.2-mm clumps that were associated with only 6.7-GHz methanol maser emission. Comparison between the 6.7-GHz methanol masers with detected 12.2-GHz with those with no detectable 12.2-GHz maser emission, showed that 12.2-GHz methanol masers were preferentially found at those 6.7-GHz methanol maser sites with associated OH emission (to a better than 95 per cent confidence). As OH masers are expected to trace a generally later stage of star formation \\citep{FC89,Cas97} than 6.7-GHz methanol masers. As the sample size of the \\citet{Lewis07} observations was small, was subject to a 6.7-GHz flux density bias and had relatively poor sensitivity, the aim of the current work was to confirm these results by testing a large sample with a much lower 12.2-GHz detection limit. Here we present a targeted search for 12.2-GHz methanol masers, with a greatly increased sample size and sensitivity (c.f. \\citet{Lewis07}), towards 113 known 6.7-GHz methanol masers which have been previously targeted for 1.2-mm dust continuum emission \\citep{Hill05}. Statistical analysis of the 1.2-mm dust clumps devoid of either methanol maser transition and those associated with 12.2-GHz and/or 6.7-GHz methanol maser sources has been carried out, with a particular emphasis on the investigation of the relative evolutionary stage each class of source is tracing. ", "conclusions": "We have searched 113 known 6.7-GHz methanol masers for associated 12.2-GHz methanol maser emission with the ATNF Parkes 64-m radio telescope. We have detected 12.2-GHz methanol masers towards 68 of these sources; a detection rate of 60 percent. We estimate the life-time of 12.2-GHz methanol maser emission to be of the order of 2.1 $\\times$ 10$^{4}$ years. We find that for the majority of sources, the peak velocity of the 12.2-GHz methanol maser emission is the same as the 6.7-GHz methanol maser and, for the remainder, the 12.2-GHz peak emission is coincident with a secondary feature present in the 6.7-GHz methanol maser spectrum. Comparison between the peak flux densities of the respective methanol maser transitions has shown that median ratio of 6.7- to 12.2-GHz peak flux density is 1:5.9, similar to previous searches. We additionally find that the ratio of 6.7- to 12.2-GHz methanol maser peak flux density has some dependence on 6.7-GHz methanol maser luminosity, with the more luminous 6.7-GHz sources having comparatively weaker 12.2-GHz methanol maser emission and vice versa. All of the target sources have previously been observed at 1.2-mm for dust continuum emission \\citep{Hill05}. We have carried out statistical analysis of the dust clump properties of sources in the following four subgroups compared the 1.2-mm dust clump properties: dust clumps with and without associated 6.7-GHz methanol maser emission; dust clumps associated with highly luminous 6.7-GHz methanol masers compared with those of low luminosity; dust clumps associated with 6.7-GHz methanol maser emission with and without associated 12.2-GHz methanol maser emission; and dust clumps with and without associated radio continuum. Our statistics show that there are many differences between the sources that fall into these subgroups; consistent with the different source combinations and characteristics being associated with different stages of massive star formation. An evolutionary scenario for the common maser species associated with high-mass star formation regions is presented. The positions of the 6.7-GHz methanol masers have been compared to sources in the GLIMPSE point source catalogue. We find that 57 per cent of the 6.7-GHz methanol masers that lie within the GLIMPSE regions are associated with a GLIMPSE point source. Comparison between the GLIMPSE colours associated with sources that exhibit emission at both 6.7- and 12.2-GHz and those exclusively at 6.7-GHz has shown that there is little difference between the colours of the associated point sources. We will present detailed investigation of the relative intensities and velocity ranges of 12.2-GHz methanol masers detected towards the 6.7-GHz methanol masers detected in the southern hemisphere component of the MMB survey \\citep{Green08} in a subsequent paper. The near simultaneous observations of hundreds of 6.7- and 12.2-GHz methanol masers collected from the MMB survey and follow-up observations will be used to make a detailed investigation of the range and distribution of the intensity ratio for these two transitions. Combining this with our evolutionary studies we will obtain unique insight into the changes in physical conditions responsible for the presence/absence of the different maser methanol transitions. Our results relating to the evolution of star formation regions will be independently tested in the future through comparison between the unbiased sample of 6.7-GHz methanol masers observed in the MMB survey \\citep{Green08}, followup observations at 12.2-GHz and 870~$\\mu$m dust sources observed as part of ATLASGAL (APEX Telescope Large Area Survey of the Galaxy)\\footnote{http://www.mpifr-bonn.mpg.de/div/atlasgal/}." }, "0910/0910.1976_arXiv.txt": { "abstract": "{ We study the evolution of curvature perturbations and the cosmic microwave background (CMB) power spectrum in the presence of an hypothesized extra anisotropic stress which might arise, for example, from the dark radiation term in brane-world cosmology. We evolve the scalar modes of such perturbations before and after neutrino decoupling and analyze their effects on the CMB spectrum. A novel result of this work is that the cancellation of the neutrino and extra anisotropic stress could lead to a spectrum of residual curvature perturbations which is similar to the observed CMB power spectrum. This implies a possible additional consideration in the determination of cosmological parameters from the CMB analysis. } \\received{\\today} \t\t% \\accepted{\\today} ", "introduction": "Anisotropic stress is the traceless component in the energy-momentum tensor and it accounts for anisotropic momentum flow in the universe. Although some previous works have considered neutrino anisotropic stress and the anisotropic stress of the primordial magnetic field \\cite{1995ApJ...455....7M,2008PhRvD..77f3003G,2006CQGra..23.4991G,2006PhRvD..74f3002G,2008arXiv0806.2018K,2008nuco.confE.226K}, they mainly considered the post neutrino decoupling era. In this work we deduce the evolution of the neutrino anisotropic stress and curvature perturbations both before and after neutrino decoupling which can be important. In the standard cosmological model, the anisotropic stress is absent before the epoch of neutrino decoupling because rapid interactions among elementary particles dissipate it. After the neutrinos have decoupled from the cosmic expansion, however, the neutrinos become relativistic free particles and neutrino anisotropic stress can grow gradually. The neutrino anisotropic stress plays an important role in the formation of large scale structure and the CMB fluctuation spectrum. In this paper, however, we consider other possible sources for anisotropic stress besides neutrinos or a magnetic field which can be present before neutrino decoupling. In particular, such terms arise naturally in cosmological theories in higher dimensions. We follow the evolution of the neutrino anisotropic stress and curvature perturbations both before and after neutrino decoupling in the presence of such anisotropic stress terms and show that curvature perturbations can stay constant on superhorizon scales as does the standard adiabatic mode. However, unlike the standard case, significant evolution of curvature perturbations occurs before neutrino decoupling. We show that, for the right conditions, such curvature perturbations might even reproduce the observed CMB power spectrum implying a possible additional consideration in the extraction of cosmological parameters from the CMB analysis. One possible source of extra anisotropic stress is the so-called \"dark radiation\" term in brane-world cosmology. This term can affect strongly the CMB anisotropies \\cite{2001PhRvD..63h4009L,2002PhRvD..65j4012L,2002PhLB..532..153B,2002PhRvD..66d3521I, 2003PhRvD..68h3518I, 2006PhRvD..73f3527U}. In such brane-world cosmology, there are two correction terms in the 4D Einstein equation. One is the extrinsic curvature which introduces a quadratic term in the energy momentum tensor, but is negligible in the low energy limit. The other is from the projected Weyl tensor. The application of the five-dimensional conservation condition to the Weyl tensor requires that the energy density term varies with scale factor as $a^{-4}$ to an observer on the brane \\cite{2001PhRvD..63h4009L}. Hence, it is called the dark radiation term and remains significant throughout the radiation-dominated epoch. There is, however, no intrinsic brane equation to determine the anisotropic stress \\cite{2001PhRvD..63h4009L}. Although there have been attempts (e.g. \\cite{2003PhRvL..91v1301K}) to constuct simple models for the Weyl anisotropic stress, the solution strongly depends upon unknown physics in the bulk dimension. Hence, in what follows we take a phenomenological approach and simply adopt the plausible assumption that in the same way that inflation generates scale-free fluctuations characterized by a spectral index $P(k) \\sim k^n$ (with $n \\sim 1$), we can expect that the Weyl anisotropic stress could be characterized by a spectral index to be determined (constrained) by a fit to the observed CMB power spectrum. Moreover, this dark-radiation term is assumed to scale as $\\propto a^{-4}$ similar to the energy density and pressure as was adopted in Ref. \\cite{2002PhLB..532..153B}. Another more familiar example of extra anisotropic stress is that of a primordial magnetic field (PMF). The amplitude of the energy density $B^2/8\\pi$ and magnetic anisotropic stress $\\rho_\\gamma\\pi_B$ of the PMF again both scale as radiation density $\\propto a^{-4}$. Moreover, such a PMF should be characterized by an amplitude and spectral index, However, the contribution of a PMF to the CMB power spectrum is constrained to be small because it has a measurable effect on the observed BB mode of the CMB on small angular scales \\cite{2008PhRvD..77d3005Y, 2008PhRvD..78l3001Y}. As shown in Ref.~\\cite{2004PhRvD..70d3011L}, the tensor mode of the primordial magnetic field has two kinds of fluctuations. One is a compensated mode, which arises from the compensation of anisotropic stress, and the other is a passive mode which was generated by the anisotropic stress of the magnetic field before the epoch of neutrino decoupling. The passive tensor mode makes a large contribution to the CMB. Although Ref.~\\cite{2004PhRvD..70d3011L} also studied the vector mode, there was no analysis of the scalar mode in that work. The spectral index of a PMF is also constrained from the effects of the associated gravity waves on primordial nucleosynthesis \\cite{2000PhRvD..61d3001D}. Hence, a PMF cannot source the scale-free extra anisotropic stress of interest to this work. We only mention it as an example of a power-law anisotropic stress which also scales as $\\propto a^{-4}$. Here we show that a scale-free anisotropic stress, if present, could imply an additional consideration in the determination of cosmological parameters from the CMB. As an extreme example, in this paper we show that an extra anisotropic stress could even lead to a CMB spectrum that agrees with observation. Such a source is a bit contrived as it must be scale invariant as naturally occurs in the inflationary scenario. Thus some kind of mechanism would have to occur to produce the desired scale-invariant anisotropic stress. Our main point is that the possibility exists that such an anisotropic stress term might be present in the early universe and that it would lead to curvature perturbations which could reproduce the observed CMB power spectrum. This possibility has not previously been pointed out to our knowledge. A second point which we make here is that this extra source of anisotropic stress may produce non-Gaussian fluctuations depending upon the nature of the source of anisotropic stress. The WMAP-5yr analysis, has indicated that there is at least a possibility for non-Gaussianity although the uncertainty is large \\cite{2008arXiv0803.0547K}. The Planck mission should constrain non-Gaussianity with high precision. If non-Gaussianity were actually observed, it could suggest (e.g.~\\cite{2004PhR...402..103B}) the need for a new cosmological paradigm which allows for non-Gaussianity such as that described here. At the very least, observational limits on non-Gaussianity could place limits on the hypothesis proposed here, depending upon the source of the extra anisotropic stress. ", "conclusions": "We have found a simple solution for Eq. (\\ref{eta2}) and have shown that if there exists an extra anisotropic stress which scales as $\\rho_\\gamma\\pi_{\\rm ex}\\propto a^{-4}$, the anisotropic stress from neutrinos $\\pi_\\nu$ exactly cancels $\\pi_{\\rm ex}$. Before neutrino decoupling, the curvature perturbations grow logarithmically. After neutrino decoupling, they become constant on superhorizon scales just like the standard adiabatic mode of inflation. This is because the total anisotropic stress vanishes via a cancellation. Thus, the resultant CMB spectrum is a superposition of the primary and passive modes. As an illustration we have considered the possibility that the passive scalar mode has a scale invariant spectrum. In this case the extra anisotropic stress might even produce a power spectrum similar to the observed CMB. This suggests a possible additional consideration in the determination of cosmological parameters from the CMB. Also we note that the Gaussianity of the CMB fluctuations depends upon the source for the extra anisotropic stress. Hence, should this extra anisotropic stress be present in the observed power spectrum, it might be detectable by non-Gaussianity in the fluctuations. Future CMB observations of non-Gaussianity may, therefore, help to constrain this possibility. To summarize, our purpose here has been to suggest that such a contribution to the observed spectrum may exist. Thus, further studies on the amplitude and spectrum of the extra anisotropic stress in brane-world cosmology are warranted." }, "0910/0910.2144_arXiv.txt": { "abstract": "A new promising development in astroparticle physics is to measure the radio emission from extensive air showers. The particles in the cascade emit synchrotron radiation (30 - 90 MHz) which is detected with arrays of dipole antennas. Recent experimental efforts are discussed. ", "introduction": "An intense branch of astroparticle physics is the study of high-energy cosmic rays to reveal their origin, as well as their acceleration and propagation mechanisms \\cite{behreview,cospar06,wuerzburg}. At energies exceeding $10^{14}$~eV cosmic rays are usually studied by indirect measurements --- the investigation of extensive air showers initiated by cosmic particles in the atmosphere. Different techniques are applied, like the measurements of particle densities and energies at ground level, or the observation of \\Cerenkov and fluorescence light. An alternative technique has been recently revitalized --- the detection of radio emission (at tens of MHz) from extensive air showers at energies exceeding $10^{16}$~eV. Radio emission from air showers was experimentally discovered in 1965 at a frequency of 44 MHz \\cite{jelleynature}. The early activities in the 1960s and 1970s are summarized in \\rref{allanrev}. Only recently, fast analog-to-digital converters and modern computer technology made a clear detection of radio emission from air showers possible. LOPES, a LOFAR Prototype Station had shown that radio emission from air showers can be detected even in an environment with relatively strong radio frequency interference (RFI) \\cite{radionature}. Further investigations of the radio emission followed with LOPES \\cite{badearadio,petrovicinclined,buitinkthunder,niglfreq,nigldirect} and the CODALEMA experiment \\cite{codalema,Ardouin:2006nb}, paving the way for this new detection technique \\cite{Falcke:2008qk}. Most likely, the dominant emission mechanism of the radio waves in the atmosphere is synchrotron radiation due to the deflection of charged particles in the Earth magnetic field (geosynchrotron radiation) \\cite{huege2003,huege2005,huegefalcke}. In the frequency range of interest ($30-90$~MHz) the wavelength of the radiation is large compared to the size of the emission region: the typical thickness of the shower disc is about 1 to 2~m only. Thus, coherent emission is expected, which yields relatively strong signals (of the order of a few $\\mu$V/m MHz) at ground level. Goal of the present activities is to further push the development of radio detection to become a new, independent way to measure the properties of air showers. Ultimate goal is to derive information about the primary, shower-inducing particle from the measurements, such as the energy and mass of the particle as well as the particle direction and point of incidence. An advantage of radio detection, e.g.\\ with respect to the fluorescence technique, is that showers can be observed with almost 100\\% duty cycle. The {\\sl arrival direction} can be inferred from the arrival time of the wavefront at the antennas. The measured radio signal at a certain distance from the shower axis is a good estimator for the number of particles at shower maximum, being almost independent of the primary particle type \\cite{huege2008}. This can be used to determine the shower {\\sl energy}. The curvature of the shower front has been investigated \\cite{lafebrecurv,icrc09-lafebre}. It could be shown that the radius of curvature measured at ground level is related to the distance to the shower maximum. The depth of the shower maximum in the atmosphere \\Xmax is an important observable to determine the mass of the primary particle. The study indicates a resolution of \\Xmax from radio observations of order of $30-40$~\\gcm2, i.e.\\ comparable to the accuracy of present fluorescence detectors. ", "conclusions": "" }, "0910/0910.0836_arXiv.txt": { "abstract": "We report on the results of a \\emph{Chandra} search for evidence of triggered nuclear activity within the Cl0023+0423 four-way group merger at $z\\sim0.84$. The system consists of four interacting galaxy groups in the early stages of hierarchical cluster formation and as such, provides a unique look at the level of processing and evolution already underway in the group environment prior to cluster assembly. We present the number counts of X-ray point sources detected in a field covering the entire Cl0023 structure, as well as a cross-correlation of these sources with our extensive spectroscopic database. Both the redshift distribution and cumulative number counts of X-ray sources reveal little evidence to suggest the system contains X-ray luminous AGN in excess to what is observed in the field population. If preprocessing is underway in the Cl0023 system, our observations suggest that powerful nuclear activity is not the predominant mechanism quenching star formation and driving the evolution of Cl0023 galaxies. We speculate that this is due to a lack of sufficiently massive nuclear black holes required to power such activity, as previous observations have found a high late-type fraction among the Cl0023 population. It may be that disruptive AGN-driven outflows only become an important factor in the preprocessing of galaxy populations during a later stage in the evolution of such groups and structures, when sufficiently massive galaxies (and central black holes) have built up, but prior to hydrodynamical processes stripping them of their gas reservoirs. ", "introduction": "There is now substantial evidence that environments of intermediate density, such as galaxy groups, play an important role in the transformation of field galaxies into the passively evolving populations found in galaxy clusters. Several studies have found that group populations already exhibit reduced star formation rates (SFR; Lewis et al.~2002; Gomez et al.~2003) and high early-type fractions similar to those observed in denser environments (Zabludoff \\& Mulchaey 1998; Jeltema et al.~2007). While the physical mechanisms responsible for the preprocessing of galaxies in the group regime are still heavily debated, several recent studies have reported an overdensity of X-ray luminous Active Galactic Nuclei (AGN) on the outskirts of clusters and within the substructure surrounding unrelaxed systems (D'Elia et al.~2004; Cappelluti et al. 2005; Kocevski et al.~2009a,2009b; Gilmour et al.~2009). These observations suggest that increased nuclear activity may be triggered in such environments and that AGN-driven outflows may play a role in suppressing star formation within galaxies during cluster assembly. Indeed the increased dynamical friction within groups and their low relative velocity dispersions make them conducive to galaxy interactions which can trigger such activity (Hickson 1997; Canalizo \\& Stockton 2001) and recent hydrodynamical simulations suggest that merger-triggered AGN feedback can have a profound effect on the gas content and star formation activity of their host galaxies (Hopkins et al.~2007; Somerville et al.~2008). Since the mass density of virialized structures increases with redshift, mergers are expected to play an even greater role in the group environment in the past. Therefore, if galaxy interactions and subsequent AGN feedback are driving a significant portion of the preprocessing found in intermediate density environments, we may expect to find an overdensity of AGN in high redshift groups in the early stages of hierarchical cluster formation. In this Letter we report on \\emph{Chandra} observations of one such system, the Cl0023+0423 (hereafter Cl0023) four-way group merger at $z=0.84$ (Lubin et al.~2009a). The Cl0023 structure consists of four interacting galaxy groups which simulations suggest are the direct progenitors of a future massive cluster. As such, the system provides a unique look at the level of processing and evolution already underway in the group environment prior to cluster assembly. To search for evidence of triggered nuclear activity within the Cl0023 structure, we present the number counts of X-ray point sources detected in a field covering the entire system, as well as a cross-correlation of these sources with our extensive spectroscopic database. Surprisingly, we find no evidence for an overdensity of X-ray detected point sources in the direction of the Cl0023 groups. We discuss the implications of this finding on the role of AGN feedback in regulating galaxy evolution in such structures. We also examine possible explanations for the lack of increased nuclear activity in the system. Throughout this Letter we assume a $\\Lambda$CDM cosmology with $\\Omega_{m} = 0.3$, $\\Omega_{\\Lambda} = 0.7$, and $H_{0} = 70$ $h_{70}$ km s$^{-1}$ Mpc$^{-1}$. \\begin{figure*} \\epsscale{1.1} \\plotone{./f1.astroph.eps} \\caption{(\\emph{left}) Adaptively smoothed density map of color-selected red galaxies in the Cl0023 field. Three density peaks that correspond with four spectroscopically confirmed galaxy groups are marked. Adapted from Lubin et al.~(2009a). (\\emph{right}) Adaptively smoothed, ACIS-I image of the Cl0023 field in the soft X-ray band (0.5-2 keV). \\label{fig-densmap}} \\end{figure*} \\vspace{0.5in} ", "conclusions": "Using \\emph{Chandra} imaging of the Cl0023 complex, we have searched for evidence of triggered nuclear activity within a dynamically active system of four galaxy groups in the early stages of cluster formation. Both the redshift distribution and cumulative number counts of X-ray point sources in the Cl0023 field reveal little evidence to suggest that the system contains X-ray luminous AGN in excess to what is observed in the field population. These results are at odds with previous reports of source excesses on the outskirts of dynamically unrelaxed clusters at high redshift. They also appear to challange the notion that AGN-driven outflows play a significant role in the preprocessing observed in galaxy groups and environments of moderate overdensity relative to the field. If preprocessing is underway in the Cl0023 system, our observations suggest that powerful (quasar mode) nuclear activity is not the predominant mechanism quenching star formation and driving the evolution of Cl0023 galaxies. Of course we cannot rule out a population of low-luminosity AGN powering ``radio mode'' feedback (Croton et al.~2006) in the Cl0023 complex as our observations are only sensitive to moderate luminosity Seyferts and QSOs. We are currently analyzing \\emph{Very Large Array} (\\emph{VLA}) 20-cm observations of Cl0023 to search for such a population and expect to present a full radio study of the system in a forthcoming paper (L.~M.~Lubin et al.~2009b, in preparation). Our current findings are in stark contrast to the overdensity of AGN recently detected in similar \\emph{Chandra} observations of the Cl1604 supercluster at $z=0.9$, where we find a population of Seyferts associated with an unrelaxed cluster and two rich groups (Kocevski et al.~2009a, 2009b). However, the galaxy populations of these groups differ in significant ways from those of the Cl0023 system. The Cl1604 groups tend to have higher velocity dispersions and more evolved galaxy populations than the Cl0023 groups, as indicated by their average SFRs and morphological fractions (Gal et al.~2008; Lubin et al.~2009a). Previous observations of Cl0023 galaxies found them to be predominately late-type systems (75\\%; Lubin et al.~1998) with substantial amounts of ongoing star formation\\footnote{This is consistent with the galaxy properties of high redshift groups with similar velocity dispersions (e.g.~Poggianti et al.~2006)} (Postman, Lubin, Oke 1998; Lubin et al.~2009a), whereas the hosts of the Cl1604 AGN tend be bulge-dominated, post-starburst galaxies which show signs of recent or ongoing galaxy interactions. Therefore, while Cl0023 contains galaxies which have the gas necessary to fuel nuclear activity, it apparently lacks the bulge-dominated and massive early-type hosts in which powerful AGN have been shown to reside (Kauffmann et al.~2003). A likely explanation for the absence of luminous AGN in the Cl0023 groups is that the system lacks galaxies with sufficiently massive nuclear black holes required to power such activity. It has previously been shown that the bulge-dominated S0 population in clusters and groups builds up over time at the expense of the spiral population and that this morphological evolution is more pronounced in lower mass systems (Poggianti et al.~2009). There is also evidence that these galaxies are typically more massive than their suspected progenitors (Dressler et al.~2009), suggesting they experience growth in their stellar bulges while in overdense environments, possibly via a centrally concentrated burst of star formation (Dressler et al.~1999). Given the correlation between bulge mass and central black hole mass (Gebhardt et al.~2000), we would expect similar growth in galactic nuclei over the same period. Therefore, if disruptive AGN-driven outflows play a role in quenching star formation in groups, as has been suggested, it may only become an important factor in the preprocessing of galaxy populations during a later stage in the evolution of such groups and structures, when sufficiently massive galaxies (and nuclear black holes) have built up, but prior to hydrodynamical processes within clusters stripping them of their gas reservoirs. Further observations of a larger sample of systems in the early stages of cluster formation, with a variety of velocity dispersions and morphological fractions, will be required to test this scenario. In the mean time, we are planning additional spectroscopic follow-up of the Cl0023 groups targeting the radio bright population as well as the remaining X-ray point sources that currently lack redshifts. This will give us a greater spectroscopic completeness of X-ray luminous AGN in the Cl0023 field, which will enable us to test our current findings and should allow us to better discern the prevalence of powerful nuclear activity during cluster formation." }, "0910/0910.2234_arXiv.txt": { "abstract": "We present a physical model for the origin of \\zsim 2 Dust-Obscured Galaxies (DOGs), a class of high-redshift ULIRGs selected at 24 \\micron \\ which are particularly optically faint ($F_{\\rm 24\\mu m}/F_R>1000$). By combining $N$-body/SPH simulations of high redshift galaxy evolution with 3D polychromatic dust radiative transfer models, we find that luminous DOGs (with \\stf$\\ga 0.3 $ mJy at $z\\sim 2$) are well-modeled as extreme gas-rich mergers in massive ($\\sim 5\\times10^{12}-10^{13}$ \\msunend) halos, with elevated star formation rates ($\\sim 500-1000$ \\msunyrend) and/or significant AGN growth ($\\dot{M}_{\\rm BH} \\ga 0.5$ \\msunyrend), whereas less luminous DOGs are more diverse in nature. At final coalescence, merger-driven DOGs transition from being starburst dominated to AGN dominated, evolving from a ``bump'' to a power-law shaped mid-IR (IRAC) spectral energy distribution (SED). After the DOG phase, the galaxy settles back to exhibiting a ``bump'' SED with bluer colors and lower star formation rates. While canonically power-law galaxies are associated with being AGN-dominated, we find that the power-law mid-IR SED can owe both to direct AGN contribution, as well as to a heavily dust obscured stellar bump at times that the galaxy is starburst dominated. Thus power-law galaxies can be either starburst or AGN dominated. Less luminous DOGs can be well-represented either by mergers, or by massive ($M_{\\rm baryon} \\approx 5\\times10^{11}$\\msun) secularly evolving gas-rich disc galaxies (with SFR $\\ga 50$ \\msunyrend). By utilising similar models as those employed in the SMG formation study of Narayanan et al. (2010), we investigate the connection between DOGs and SMGs. We find that the most heavily star-forming merger driven DOGs can be selected as Submillimetre Galaxies (SMGs), while both merger-driven and secularly evolving DOGs typically satisfy the \\bzk \\ selection criteria. The model SEDs from the simulated galaxies match observed data reasonably well, though Mrk 231 and Arp 220 templates provide worse matches. Our models provide testable predictions of the physical masses, dust temperatures, CO line widths and location on the \\magorrian \\ relation of DOGs. Finally, we provide public SED templates derived from these simulations. ", "introduction": "\\label{section:introduction} Redshift \\zsim 2 is a rich epoch for understanding galaxy formation and evolution. During this time period, the bulk of the cosmic stellar mass was assembled \\citep{dic03,rud06}, and the star formation rate and quasar space densities are both near their peaks \\citep{bou04,hop04,hop07,sha96}. In the local Universe, infrared (IR) selection of galaxies is an efficient means of selecting galaxies undergoing significant star formation (e.g., Arp 220), and possibly hosting optically-obscured active galactic nuclei (AGN; e.g., Mrk 231, NGC 6240). By analogy, at higher redshifts, IR selection techniques select galaxies which are major contributors to the cosmic star formation rate density, far-IR background, and progenitors of the present-day massive galaxy population. Indeed, ultraluminous infrared galaxies (ULIRGs; \\lir $>$ 10$^{12}$ \\lsunend) at \\zsim 2 appear to contribute substantially to the infrared luminosity density \\citep[]{per05,cap07,red08,hop10,hop10b}, and the \\z$>$0.7 star formation rate density \\citep{lef05,shi09}. A well-studied sample of galaxies selected for their FIR properties at \\zsim 2 are Submillimetre Galaxies (SMGs), chosen in blank field surveys for their redshifted cool dust emission at 850 \\micron \\ above \\sef $\\ga$ 5 mJy. These galaxies are in a transient starburst phase \\citep[e.g.,] []{swi04,men07,you08b,mic09}, host rapidly growing supermassive black holes \\citep[e.g.,] []{ale05a,ale08}, and reside in extremely massive $\\sim$5$\\times$10$^{12}$ \\msun halos \\citep{bla04}. While these galaxies represent an important sub-population of infrared-luminous galaxies at the epoch of peak galaxy formation, there is some concern that a selection at 850 \\micron \\ may be missing significant populations of high-redshift ULIRGs with warmer dust temperatures \\citep[e.g.,] []{bus09,cha09,cas09,hua09,you09b}. As such, alternative selection techniques for identifying \\zsim 2 ULIRGs have become desirable. With the launch of the {\\it Spitzer Space Telescope}, surveys at 24 \\micron \\ have uncovered a population of ULIRGs at \\zsim 2 \\cite[e.g.,] [ and references therein]{rig04,don07,yan07,far08,soi08,lon09,hua09}. Recently, \\citet[][ hereafter, D08; see also \\citet{fio08}]{dey08} presented an efficient means of identifying high-redshift ULIRGs which are both mid-infrared luminous and optically faint. Specifically, by imposing a nominal selection criteria of $R$-[24] $>$ 14 (roughly corresponding to a flux density ratio \\stf/\\sr $>$ 1000; formally \\stf/\\sr $> 960$) and \\stf $>$ 0.3 mJy, D08 found a sample of ULIRGs with a relatively narrow redshift distribution (centered at $<$\\z$>$=1.99 with $\\sigma_z$=0.5) which exhibit heavy reddening of the intrinsic UV flux and strong rest-frame 8 \\micron \\ emission \\citep{hou05,fio08}. D08 refer to these galaxies as Dust-Obscured Galaxies (DOGs)\\footnote{We note that while this paper nominally focuses on 24 \\micron-selected galaxies which additionally have \\tfr $> 1000$, the results of this paper are generally applicable to the 24 \\micron-bright (\\stf $> 0.3 $mJy) \\zsim 2 ULIRG population.}. DOGs exhibit a range of luminosities, though AGN-dominated galaxies tend to preferentially dominate the bright end of the 24 \\micron \\ flux density distribution \\citep{bra06,wee06b,wee06a,dey08,fio08,fio09,sac09}. The DOGs with the highest 24 \\micron \\ flux densities (\\stf $>$ 0.8 mJy) have bolometric infrared luminosities (\\lir) $\\sim$ 10$^{13}$ \\lsun \\citep{bus09b,tyl09}. As a population, the subset of \\stf$>0.3$ mJy galaxies which are also DOGs contribute around a quarter of the total infrared luminosity density at \\zsim 2 \\citep{dey08}, and constitute $\\sim$60\\% of the total contribution by \\zsim 2 ULIRGs. It is clear that DOGs are a cosmologically significant population of galaxies with diverse observational characteristics. Broadly, 24 \\micron-selected ULIRGs at \\zsim 2 (including DOGs) appear to fall into two categories based on their mid-IR IRAC SEDs: those with a power-law shape and those which exhibit a bump at observed $\\lambda \\sim 5$ \\micron \\ corresponding to the 1.6 \\micron \\ rest frame stellar photospheric bump (hereafter, these are referred to as PL galaxies and bump galaxies, respectively). Bump galaxies are thought to be dominated by star formation, with the bump (at rest-frame 1.6 \\micron) originating from starlight. PL galaxies are thought to have their mid-IR SED dominated by AGN continuum emission \\citep{wee06a,don07,yan07,mur09}. In support of this scenario, PL galaxies have SEDs which are reasonably represented by that of a Mrk 231 template \\citep{bus09b,tyl09}, and tend to show relatively compact optical morphologies \\citep[$R_e\\sim 1-6$ kpc; ][]{mel08,mel09,bus09}. On the other hand, bump galaxies are typically polycyclic aromatic hydrocarbon (PAH)-rich, show PAH equivalent widths consistent with being star formation dominated \\citep{yan05,saj07,far08,des09,hua09}, and are more often morphologically resolved into multiple components \\citep{bus09}. However, \\citet{fio08,fio09} have argued that even these less luminous galaxies may harbor heavily obscured AGN. DOGs are massive galaxies, with stellar masses $M_\\star >$ 10$^{10-11}$ \\msun \\citep{bus09,lon09,hua09}, and cluster in group-sized halos of order \\citep[$M_{\\rm DM} \\approx 10^{12-13}$ \\msunend;][]{bro08}. The clustering is luminosity-dependent, with more luminous DOGs inhabiting more massive halos. Since the more luminous DOGs also tend to exhibit power-law SEDs, it follows that PL DOGs cluster more strongly than the lower luminosity bump DOGs. Although much progress has been made in understanding DOGs in a relatively short time period, a myriad of fundamental questions regarding their physical nature exist. Are DOGs preferentially isolated galaxies or mergers (or both)? Does the \\tfr \\ criterion select galaxies at a particular evolutionary point? How are bump DOGs and PL DOGs related - are they distinct galaxy populations, or possibly related via an evolutionary sequence? How are DOGs related to other coeval high-redshift populations (e.g., the SMGs, \\bzk \\ galaxies, etc.)? Numerical simulations can offer complementary information to the observations in hand, and provide insight into the aforementioned questions. In particular, by coupling radiative transfer modeling with hydrodynamic simulations of galaxy evolution, one can precisely mimic local and high-redshift observational selection functions, and directly relate observed trends to physical conditions in the model galaxies \\citep[e.g.,] []{jon06a,jon06b,jon10,lot08,nar10a,you09}. Previously, we have utilised similar methods to investigate the formation and evolution of high-redshift Submillimetre Galaxies \\citet[][C. Hayward et al. in prep.]{nar09b,nar10a}. Here, we build upon these efforts by utilising similar models as those employed in \\citet{nar09b} and \\citet{nar10a}. This will allow us not only to investigate a potential formation mechanism for DOGs, but to explore the potential connection between \\zsim 2 ULIRGs selected for their 24 \\micron \\ brightness, and those selected in the submillimetre. The paper is organised as follows. In \\S~\\ref{section:methods}, we detail our numerical methods. In \\S~\\ref{section:dogformation}, we describe the formation of \\zsim 2 DOGs, and explain the evolution of the 24/R ratio and the relationship between bump and PL DOGS. In \\S~\\ref{section:physicalnature}, we discuss the physical requirements necessary to form a DOG, and whether mergers are necessary; In \\S~\\ref{section:othergalaxies}, we detail the relationship between DOGs, SMGs and \\bzk \\ galaxies. In \\S~\\ref{section:observations}, we relate our model to existing observations, and make testable predictions. In \\S~\\ref{section:discussion}, we provide discussion. We conclude in \\S~\\ref{section:conclusions}. Throughout this paper, we utilise a fiducial DOG selection criteria \\tfr \\ $>$ 1000 and \\stf $>$ 300 \\microjy, consistent with the DOGs sample in D08. We note again that while we focus on these 24 \\micron \\ selected galaxies which are optically faint (DOGs), the results are generally applicable to most 24 \\micron -selected \\zsim 2 ULIRGs. Throughout, we assume a cosmology with $\\Omega_\\Lambda=0.7$, $\\Omega_m=0.3$ and $h=0.7$. Finally, we note that the models presented here are not cosmological in nature. As such, all non-evolutionary plots should be taken to represent ranges of expected values and colors, rather than true expected distributions. \\begin{table*} \\label{table:ICs} \\centering \\begin{minipage}{100mm} \\caption{DOGs are ordered with decreasing halo mass (and grouped by mergers and isolated galaxies, such that the isolated galaxies have the prefix 'i' in their name). Column 1 is the name of the model used in this work. We emphasise that these models are nearly identical to those employed in the SMG formation study by \\citet{nar10a} and \\citet{nar09b}. This will facilitate the comparison of these models with SMGs (\\S~\\ref{section:smg}). Columns 2 and 3 give the virial velocity and halo mass of the galaxies. Column 4 is the total baryonic mass of the system. Column 5 is the merger mass ratio. Columns 6 \\& 7 are initial orientations for disc 1, Columns 8 \\& 9 are for disc 2. The orientation angles are with respect to the merger plane (such that $\\theta , \\phi = 0$ would be a coplanar merger). Column 10 lists the initial gas fraction of the simulation.} \\begin{tabular}{@{}cccccccccc@{}} \\hline Model & V$_{\\rm c}$ & $M_{\\rm DM}$ & $M_{\\rm bar}$& Mass Ratio & $\\theta_1$&$\\phi_1$ & $\\theta_2$ & $\\phi_2$ &$f_g$\\\\ &(\\kmsend)&\\msun&\\msun&&&&\\\\ \\hline DOG1 & 500:500 & 1.25$\\times$10$^{13}$:1.25$\\times$10$^{13}$ & 1.1$\\times$10$^{12}$& 1:1 & 30 & 60 & -30 & 45 &0.8,0.6,0.4\\\\ DOG2 & 500:500 & 1.25$\\times$10$^{13}$:1.25$\\times$10$^{13}$ & 1.1$\\times$10$^{12}$& 1:1& 360 & 60 & 150 & 0& 0.8\\\\ DOG3 & 500:500 & 1.25$\\times$10$^{13}$:1.25$\\times$10$^{13}$& 1.1$\\times$10$^{12}$& 1:1 &-109 & -30 & 71 &-30& 0.8\\\\ DOG4 & 500:320 & 1.25$\\times$10$^{13}$:3.4$\\times$10$^{12}$ & 6.9$\\times$10$^{11}$&1:3 & 30 & 60 & -30 & 45& 0.8,0.6,0.4\\\\ DOG5 & 500:320 & 1.25$\\times$10$^{13}$:3.4$\\times$10$^{12}$& 6.9$\\times$10$^{11}$&1:3 & 360 & 60 & 150 & 0 & 0.8\\\\ DOG6 & 500:320 & 1.25$\\times$10$^{13}$:3.4$\\times$10$^{12}$& 6.9$\\times$10$^{11}$&1:3& -109 & -30 & 71 & -30 & 0.8\\\\ DOG7 & 500:225 & 1.25$\\times$10$^{13}$:1.2$\\times$10$^{12}$ & 6.0$\\times$10$^{11}$&1:10 & 30 & 60 & -30 &45 & 0.8,0.6,0.4\\\\ DOG8 & 500:225 & 1.25$\\times$10$^{13}$:1.2$\\times$10$^{12}$ & 6.0$\\times$10$^{11}$&1:10 & 360 & 60 & 150 & 0 & 0.8\\\\ DOG9 & 500:225 & 1.25$\\times$10$^{13}$:1.2$\\times$10$^{12}$ & 6.0$\\times$10$^{11}$&1:10 & -109 & -30 & 71 &-30 & 0.8\\\\ DOG10 & 320:320 & 3.4$\\times$10$^{12}$:3.4$\\times$10$^{12}$ &2.9$\\times$10$^{11}$ & 1:1 & 30 & 60 & -30 &45 &0.8,0.6,0.4\\\\ DOG11 & 320:320 & 3.4$\\times$10$^{12}$:3.4$\\times$10$^{12}$ &2.9$\\times$10$^{11}$ & 1:1 & 360 & 60 & 150 &0 &0.8\\\\ DOG12 & 320:320 & 3.4$\\times$10$^{12}$:3.4$\\times$10$^{12}$ &2.9$\\times$10$^{11}$ & 1:1 &-109 & -30 & 71 &-30& 0.8\\\\ DOG13 & 225:225 & 1.2$\\times$10$^{12}$:1.2$\\times$10$^{12}$ & 1.0$\\times$10$^{11}$& 1:1 & 30 & 60 & -30 &45 &0.8,0.6,0.4\\\\ iDOG1 & 500 & 1.25$\\times$10$^{13}$ &5.5$\\times$10$^{11}$& --- & N/A & N/A&N/A & N/A& 0.8,0.6,0.4\\\\ iDOG2 & 320 & 3.4$\\times$10$^{12}$ & 1.5$\\times$10$^{11}$& --- & N/A & N/A&N/A & N/A &0.8,0.6,0.4\\\\ iDOG3 & 225 & 1.2$\\times$10$^{12}$ &5$\\times$10$^{10}$& ---& N/A &N/A& N/A & N/A &0.8,0.6,0.4\\\\ \\hline \\end{tabular} \\end{minipage} \\end{table*} ", "conclusions": "\\label{section:conclusions} Utilising a combination of polychromatic radiative transfer calculations and hydrodynamic simulations of high-redshift galaxy evolution, we have formulated a physical model for the formation and evolution of \\zsim 2 Dust-Obscured Galaxies. We have utilised the very similar models as those employed in previous studies aimed at investigating the formation and evolution of a similar class of \\zsim 2 ULIRGS: the Submillimetre Galaxy population. While the uncertainty in the models is dominated by small-scale physics which is unresolved by the simulations, our methodology allows us to make the following general conclusions regarding the origin of DOGs, and their relationship to other high-\\z \\ ULIRGs: \\begin{itemize} \\item DOGs are a diverse class of galaxies, ranging from secularly evolving star-forming disc galaxies (forming stars at $\\sim$50-100 \\msunyrend) to extreme gas-rich galaxy mergers forming stars at $\\ga$ 1000 \\msunyrend. \\item The most luminous DOGs (\\stf $\\ga$ 300 \\microjy) are well represented by mergers of massive ($M_{\\rm DM} \\approx 5 \\times 10^{12}-10^{13}$ \\msunend) galaxies. These galaxies are actively transitioning from being starburst dominated to being AGN dominated. At decreasing 24 \\micron \\ flux densities (\\stf $\\la$ 100 \\microjy), DOGs may either be galaxy mergers, or secularly evolving disc galaxies with more modest ($\\sim$50-100 \\msunyrend) SFRs. \\item Luminous, merger-driven DOGs naturally transition from having a bump-like mid-IR SED to a power-law (PL) shape, as the dominant power source transitions from star formation to the central AGN. That said, there is an overlap period where star-formation dominated sources can appear as PL galaxies. \\item The most luminous DOGs assemble both significant stellar masses ($M_\\star \\approx 10^{11}$ \\msunend), as well as contribute toward the growth of supermassive ($M_{\\rm BH} \\approx 10^9$ \\msunend) black holes. \\item Merger-driven DOGs overlap with the \\zsim 2 Submillimetre Galaxy population. SMGs generally represent the earlier, starburst-dominated phase of the DOGs evolution. This can be tested via the location of SMGs and DOGs on the \\magorrian \\ relation. \\end{itemize} In advance of {\\it Herschel}, {\\it JWST} and ALMA, we provide the following testable predictions for our model of DOG formation: \\begin{itemize} \\item We quantify the expected range of black hole, \\htwo, stellar, and halo masses for luminous (\\stf $>$ 300 \\microjy) DOGs as well as SMGs. \\item We provide dust temperatures for DOGs selected at various 24 \\micron \\ flux density limits. \\item We detail the location of SMGs and bright (\\stf $>$ 1 mJy) DOGs on the \\magorrian \\ relation as a test for the modeled evolutionary scenario that SMGs evolve into PL DOGs. \\item We provide a prediction for the CO line widths from DOGs, and suggest that they will be of order the largest observed line widths at high redshift, comparable to \\zsim 2 SMGs. \\end{itemize} Finally, we provide model SED templates for \\z=2 DOGs as a function of total infrared luminosity (\\lir = 1-1000 \\micron). These are provided publicly, along with an IDL script for easily reading in the templates at http://www.cfa.harvard.edu/~dnarayan/DOG\\_SED\\_Templates.html" }, "0910/0910.4718_arXiv.txt": { "abstract": "The Array for Microwave Background Anisotropy (AMiBA) is a radio interferometer for research in cosmology, currently operating 7 0.6m diameter antennas co-mounted on a 6m diameter platform driven by a hexapod mount. AMiBA is currently the largest hexapod telescope. We briefly summarize the hexapod operation with the current pointing error model. We then focus on the upcoming 13-element expansion with its potential difficulties and solutions. Photogrammetry measurements of the platform reveal deformations at a level which can affect the optical pointing and the receiver radio phase. In order to prepare for the 13-element upgrade, two optical telescopes are installed on the platform to correlate optical pointing tests. Being mounted on different locations, the residuals of the two sets of pointing errors show a characteristic phase and amplitude difference as a function of the platform deformation pattern. These results depend on the telescope's azimuth, elevation and polarization position. An analytical model for the deformation is derived in order to separate the local deformation induced error from the real hexapod pointing error. Similarly, we demonstrate that the deformation induced radio phase error can be reliably modeled and calibrated, which allows us to recover the ideal synthesized beam in amplitude and shape of up to 90\\% or more. The resulting array efficiency and its limits are discussed based on the derived errors. ", "introduction": "The Array for Microwave Background Anisotropy (AMiBA) is a forefront radio interferometer for research in cosmology. The project is led by the Academia Sinica, Institute of Astronomy and Astrophysics (ASIAA), Taiwan, with major collaborations with National Taiwan University, Physics Department (NTUP), Electrical Engineering Department (NTUEE), and the Australian Telescope National Facility (ATNF). Contributions also came from the Carnegie Mellon University and NRAO. As a dual-channel 86-102 GHz interferometer array of up to 19 elements, AMiBA is designed to have full polarization capabilities, sampling structures greater than $2^{\\prime}$ in size. The AMiBA targets the distribution of clusters of galaxies via the Sunyaev-Zel'dovich (SZ) Effect$^1$. It will also measure the Cosmic Microwave Background (CMB) temperature anisotropies$^2$ on scales which are sensitive to structure formation scenarios of the Universe. AMiBA is sited on Mauna Loa, Big Island, Hawaii, at an elevation of 3,425m, currently operating the initial phase with 7 antennas of 0.6m diameter in a compact configuration, Fig.\\ref{front_hexpol0_bw}. The correlator and the receivers with the antennas are co-mounted on a fully steerable 6m diameter platform controlled by a hexapod mount. The AMiBA hexapod mount with its local control system was designed and manufactured by Vertex Antennentechnik GmbH, Duisburg, Germany. The 0.6m diameter Cassegrain antenna$^3$ field-of-view (fov) is $\\sim 22^{\\prime}$ at Full Width Half Maximum (FWHM). The synthesized beam resolution in the hexagonally compact configuration (0.6m, 1.04m and 1.2m baselines) is $\\sim 6^{\\prime}$. This sets our target pointing accuracy to about $\\sim 0.6^{\\prime}$. \\\\ In parallel to the current 7-element observation program, AMiBA is undergoing a 13-element expansion with 1.2m diameter Cassegrain antennas (fov $\\sim 11^{\\prime}$). The resulting synthesized beam of about $\\sim 2^{\\prime}$ requires then ideally an improved pointing accuracy of about $\\sim 0.2^{\\prime}$. The platform photogrammetry tests$^{4,5}$ have revealed local deformations which might affect individual antenna pointings (including the optical telescope pointing) and their radio phases for the longer baselines. It therefore seems desirable to include such a correction in the current pointing error model and in the phase correction scheme. \\\\ The goal of this paper is to investigate a refined pointing error and phase correction model for the AMiBA expansion phase. The adopted approach can be of general interest since a platform deformation for a hexapod of this size is likely to be a generic problem depending on the level of accuracy which is needed. The paper is organized as following: Section \\ref{7element} summarizes the pointing error model used for the currently operating 7-element compact configuration. Section \\ref{13element} presents the approach taken to improve the pointing accuracy and the phase correction for the 13-element expansion and it also discusses its limitations. Previous AMiBA status reports and the current observations are described in a series of papers.$^{6,7,8,9}$ \\begin{figure} \\begin{center} \\includegraphics[scale=0.95]{front_hexpol0_archive.eps} \\caption{\\label{front_hexpol0_bw} Front view of the AMiBA. Installed are 7 antennas in compact configuration, giving 0.6m, 1.04m and 1.2m baselines. Free receiver holes in the platform allow for different array configurations and for the expansion phase with 13 antennas. The 1st 8 inch refractor (OT1) for optical pointing is attached to the black bracket below the lowermost antenna at a distance of 1.4m from the platform center. A second optical telescope (OT2) is located at 90$^{\\circ}$ to the left hand side at the same radius.} \\end{center} \\end{figure} \\begin{figure} \\begin{center} \\includegraphics[scale=0.8]{ot1_ot2_bw.eps} \\caption{\\label{ot1_ot2__bw}OT1 (in the back) and OT2 (in the front, lower left hand side corner), mounted with a $90^{\\circ}$ phase shift in between them at a radius of 1.4m. Clearly seen are also the reflecting photogrammetry targets.} \\end{center} \\end{figure} ", "conclusions": "We have demonstrated that the local platform deformation error can be isolated from the pointing error with the help of two OTs. The extracted deformation error is in the range $-1'$ to $+1'$ which is consistent with the photogrammetry data. The verification of an improved pointing accuracy is, however, non-trivial, because the platform deformation component will always be present on the ccd images. Therefore, one always has to rely on the modeling. Aiming for a pointing accuracy of $\\sim 0.5'$ rms ($\\eta_{abs}\\sim 0.76$) seems reasonable. Even smaller pointing errors improve the efficiency only marginally. Similarly, the platform deformation induced phase error is corrected with the photogrammetry data, and the synthesized beam amplitude is restored to 90\\% or more. \\\\ Finally, one has to ask whether the separation of platform and pointing error is needed at all. Assuming that the errors are not separated and the derived pointing errors are then reduced to zero, one still ends up with an effective pointing error ($p_{az}$ and $p_{el}$ in Eq.(\\ref{delta_abs})) in the range $[-1,1]$ due to the over-/under-compensated platform deformation error. $\\delta_{abs}$ can therefore not be further reduced and the efficiency is limited to 67\\%. For large pointing errors ($>1'$) the separation becomes unnecessary. Going beyond $\\eta_{abs}\\sim 0.67$ requires a better effective pointing and a separation of the error components. In any case, a phase correction model is needed to restore the synthesized beam." }, "0910/0910.4835_arXiv.txt": { "abstract": "{The e-VLBI technique offers a unique opportunity for users to probe the milliarcsecond (mas) scale structure of unidentified radio sources, and organise quick follow-up observations in case of detection. Here we report on e-EVN results for a peculiar radio source that has been suggested to act as a gravitational lens. However the lensing galaxy has not been identified in the optical or the IR bands so far. Our goal was to look for an active galactic nucleus (AGN) in this suspected dark lens system. The results indicate strong AGN activity, and rule out the possibility that the radio source itself is gravitationally lensed.\\ } \\FullConference{Science and Technology of Long Baseline Real-Time Interferometry: \\\\ The 8th International e-VLBI Workshop, EXPReS09\\\\ June 22 - 26 2009\\\\ Madrid, Spain} \\begin{document} ", "introduction": " ", "conclusions": "" }, "0910/0910.3412_arXiv.txt": { "abstract": "{ We present synthetic broad-band photometric colors of a late-type giant located close to the RGB tip ($T_{\\rm eff}\\approx3640$\\,K, $\\log g=1.0$ and ${\\rm [M/H]}=0.0$). Johnson-Cousins-Glass \\emph{BVRIJHK} colors were obtained from the spectral energy distributions calculated using 3D hydrodynamical and 1D classical stellar atmosphere models. The differences between photometric magnitudes and colors predicted by the two types of models are significant, especially at optical wavelengths where they may reach, e.g., $\\Delta V\\approx0.16$, $\\Delta R\\approx0.13$ and $\\Delta (V-I)\\approx0.14$, $\\Delta (V-K)\\approx0.20$. Differences in the near-infrared are smaller but still non-negligible (e.g., $\\Delta K\\approx 0.04$). Such discrepancies may lead to noticeably different photometric parameters when these are inferred from photometry (e.g., effective temperature will change by $\\Delta T_{\\rm eff}\\approx60$\\,K due to difference of $\\Delta (V-K)\\approx0.20$). ", "introduction": "Late-type giants are important tracers of intermediate age and old stellar populations in the Galaxy and beyond. Thus, the availability of reliable stellar atmosphere models is of utmost importance for understanding the structure and evolution of these stars and their host populations. The advent of the 3D hydrodynamical codes allowed to accommodate a more realistic treatment of non-stationary phenomena (e.g., convection) in the stellar atmosphere modeling and thus it is natural to expect that new 3D hydrodynamical models may provide important insights about observable properties and the interior structures of late-type giants. Indeed, recent work of \\citet{CAT07}, \\citet{C08} demonstrates that there are significant differences in the spectral line strengths predicted by the 3D hydrodynamical and 1D classical stellar atmosphere models. This leads to substantial discrepancies between the elemental abundances of various chemical species derived with the classical 1D and 3D hydrodynamical stellar atmosphere models \\citep[see][for details]{CAT07}. Still, the extent of differences in the global spectral properties predicted by the 1D classical and 3D hydrodynamical models is largely unknown. In one of our previous studies we have used a simplified approach to show that significant differences can be expected in the photometric colors predicted by the two types of models \\citep{KHL05}. Whether these conclusions would also hold with the photometric colors based on the results of full 3D spectral synthesis still had to be confirmed. In this study we extend the previous work and calculate synthetic photometric colors of a late-type giant from the full 3D spectral energy distributions. For this purpose we use a model of a late-type giant which is significantly cooler than those considered by Collet et al. (2007), with its atmospheric properties typical to those of solar-metallicity late-type giants located close to the RGB tip. \\begin{figure*}[t!] \\begin{centering} \\resizebox{10cm}{!}{\\includegraphics[clip=true]{kucinskas_fig02.eps}} \\caption{\\footnotesize Differences between the photometric magnitudes of a late-type giant predicted by 3D hydrodynamical and 1D classical models: ${\\rm 3D}-{\\rm 1D}$ (black solid line) and ${\\rm 3D}-\\mD$ (red/gray dashed line).} \\label{colors-BVRIJHK} \\end{centering} \\end{figure*} ", "conclusions": "The results obtained allow us to conclude that convection indeed plays an important role in defining the intrinsic atmospheric structures and observed properties of late-type giants. Specifically, we find that spectral energy distributions and photometric colors of a late-type giant produced with 3D hydrodynamical and 1D classical stellar atmosphere models are substantially different. Differences in photometric magnitudes and colors are considerably larger than typical photometric errors (e.g., $\\Delta V\\approx0.16$, $\\Delta (V-K)\\approx0.20$). These differences may result in further discrepancies, for instance, in the photospheric parameters derived from photometric colors (e.g., a difference of $\\Delta (V-K)\\approx0.20$ will change the estimated effective temperature by $\\Delta T_{\\rm eff}\\approx60$\\,K). Obviously, this may have direct consequences to any photometric work that relates to late-type giants and thus once again stresses the importance of 3D hydrodynamical model atmospheres to be used in the interpretation of observational data." }, "0910/0910.4697_arXiv.txt": { "abstract": "We develop an algorithm of separating the $E$ and $B$ modes of the CMB polarization from the noisy and discretized maps of Stokes parameter $Q$ and $U$ in a finite area. A key step of the algorithm is to take a wavelet-Galerkin discretization of the differential relation between the $E$, $B$ and $Q$, $U$ fields. This discretization allows derivative operator to be represented by a matrix, which is exactly diagonal in scale space, and narrowly banded in spatial space. We show that the effect of boundary can be eliminated by dropping a few DWT modes located on or nearby the boundary. This method reveals that the derivative operators will cause large errors in the $E$ and $B$ power spectra on small scales if the $Q$ and $U$ maps contain Gaussian noise. It also reveals that if the $Q$ and $U$ maps are random, these fields lead to the mixing of the $E$ and $B$ modes. Consequently, the $B$ mode will be contaminated if the powers of $E$ modes are much larger than that of $B$ modes. Nevertheless, numerical tests show that the power spectra of both $E$ and $B$ on scales larger than the finest scale by a factor of 4 and higher can reasonably be recovered, even when the power ratio of $E$- to $B$-modes is as large as about 10$^2$, and the signal-to-noise ratio is equal to 10 and higher. This is because the Galerkin discretization is free of false correlations, and keeps the contamination under control. As wavelet variables contain information of both spatial and scale spaces, the developed method is also effective to recover the spatial structures of the $E$ and $B$ mode fields. ", "introduction": "The scalar component of primordial perturbations of the universe can be detected by the maps of temperature fluctuations of Cosmic Microwave Background Radiation (CMBR), while the tensor component of the primordial perturbations has to be probed by the maps of the Stokes parameter $Q$ and $U$ of the linear polarization of the CMBR. A tensor field generally contains electric-like $E$-mode and magnetic-like $B$-modes. In the linear regime, vortical mode of primordial perturbations do not grow during the clustering of density field, and therefore, the perturbed field initially has to be curl-free. That is, the primordial perturbations can only yield the $E$-mode, but not $B$-mode of the CMBR polarization field. On the other hand, $B$-mode perturbations can be produced by gravitational waves. Therefore, extracting the $B$-mode information from CMBR polarization maps is crucial to verify the existence of gravitational wave background produced at the inflationary epoch. Moreover, gravitational lensing of clusters and hot electron scattering of reionization would be able to yield both $E$- and $B$-modes. To study these problems a sharp decomposition of $E$- and $B$-modes from $Q$ and $U$ maps is required. If both $Q$ and $U$ maps are available over the whole sky, one can find the whole sky maps of $E$- and $B$-modes with the spherical harmonic decomposition, because the relation between the maps of ($Q$, $U$) and ($E$, $B$) in the space spanned by bases of spin two harmonics is local (Kamionkowsky et al. 1997; Zaldarriaga \\& Seljak 1997). However, the observed maps cannot be global; it is always limited by the contamination of our galaxy and other foreground sources. The relation between the maps of ($Q$, $U$) and ($E$, $B$) in physical space contains the Laplace operator, and therefore, it is non-local. The $E/B$ decomposition with the spatially-limited maps of $Q$ and $U$ will not be unique if we lack of information of the polarization and its derivative on the boundary of the maps. For noiseless samples, the problem of uniqueness would be solved by constructing orthogonal modes with window functions to fit the requirements of boundary conditions (Lewis et al. 2002; Bunn et al. 2003; Smith 2006; Smith \\& Zaldarriaga 2007). It is, however, similar to the domain (or windowed) Fourier analysis. The result will not be useful to study the structures in physical space (e.g. Chiueh \\& Ma 2002). The other challenge caused by the derivative operator is because the $Q$ and $U$ maps are discrete. Mathematically, the derivative operators $\\partial_x$ or $\\partial^2_x$ are continuous linear operators mapping functions defined in Hilbert space, while the observed samples $Q$ and $U$ actually are defined in space spanned by base $v_i$, $i\\in I$, which is a set of finite indices. This difference leads to large numerical errors when the discrete maps are noisy. The last, but not least, problem is from the smallness of $B$-modes. On the scale of one degree order, the power of $B$-mode caused by gravitational waves at inflationary epoch is less than that of $E$-modes by a factor of at least 10$^2$. As the maps of $Q$ and $U$ are random fields, the variance of the random fields will lead to the mixing of $E$- and $B$-modes. Consequently, $E$- and $B$-modes would be contaminated from each other. Therefore, it is difficult to recover the power of $B$-mode if the powers of $E$ modes are much larger than that of $B$ modes. In this paper, we develop an algorithm of the $E/B$ decomposition based on the discrete wavelet transform (DWT) analysis, which is a compromise between the decompositions in physical space and scale space. The DWT analysis of the CMBR temperature fluctuation maps has attracted much attention in the last decade (Pando et al. 1998; Sanz et al. 1999; Mukherjee et al. 2000). Besides these points, we especially take the advantage of the so-called wavelet-Galerkin discretization (e.g. Louis et al. 1997), which is to approximate derivative operator to a matrix in space spanned by wavelet bases. For some available wavelets, the matrixes are exactly diagonal in scale-space, and narrowly banded in spatial space. This made the uncertainties from boundary, noises and variances are under control. We will study the conditions, under which the information of small $B$-mode can approximately be extracted from noisy maps of $Q$ and $U$. The paper is organized as follows. Section 2 presents the method of the $E/B$ separation in the DWT space. Section 3 tests the DWT algorithm with samples with known spatial structures. We show that the method effectively to suppresses the uncertainties from boundary effect and noise. It is also effective to identify spatial structures of $E$ and $B$ fields. Section 4 presents the tests for samples of Gaussian random field. The effect of the variance of Gaussian random field is analyzed, especially the problem of the mixing of $E$- and $B$-modes. Section 5 addresses the effectiveness of the wavelet-Galerkin discretization. Finally conclusions are given in Section 6. The DWT representation of derivative operators are given in the Appendix. ", "conclusions": "The algorithm developed in this paper can be summarized as follows \\begin{itemize} \\item From observed noisy and discrete maps $Q(x,y)$ and $U(x,y)$ we calculate their DWT maps $Q_{l_1,l_2}$ and $U_{l_1,l_2}$ on the finest scale ${\\bf j}=(J,J)$, which is given by the resolution. \\item Using eqs.(16) and (17) we decompose $Q_{l_1,l_2}$ and $U_{l_1,l_2}$ into $\\mathbb{E}_{l_1,l_2}$ and $\\mathbb{B}_{l_1,l_2}$. \\item Using eqs.(19) and (20) we calculate WFCs $\\tilde{\\epsilon}^E_{\\bf j,l}$ and $\\tilde{\\epsilon}^B_{\\bf j,l}$ on scales $(j_1,j_2)$ and $j_1,j_2 \\leq J$. \\item Using the WFC maps $\\tilde{\\epsilon}^E_{\\bf j,l}$ and $\\tilde{\\epsilon}^B_{\\bf j,l}$ we calculate the DWT power spectra by eqs.(30) and (31). \\item We identify spatial structures with maps of $\\mathbb{E}_{l_1,l_2}$ and $\\mathbb{B}_{l_1,l_2}$. \\end{itemize} With this algorithm, it is possible to recover the power spectrum of $B$-mode random fields from noisy Stokes parameter maps $Q$ and $U$ when the power ratio $E/B$ is as high as $10^2$, and the S/N is equal to or higher than 10. For samples with given structures, the $B$-mode structure can also be identified when the power ratio $E/B$ is equal to 10$^2$. Besides power spectrum, the DWT variables of SFCs ($\\epsilon^E_{\\bf j,l}$, $\\epsilon^B_{\\bf j,l}$) and WFCs ($\\tilde{\\epsilon}^E_{\\bf j,l}$, $\\tilde{\\epsilon}^B_{\\bf j,l}$) can be used for high order statistics, such as high order moments, scale-scale correlation, cross correlation between the $E$ and $B$ and other maps. With the DWTs, one can construct orthogonal, divergence-free vector wavelets. It has been used for a local analysis of the velocity field of incompressible turbulence (Urban 1995; Kishida et al. 1999; Albukrek et al. 2002). The divergence-free $B$ field is similar to a 2-D velocity field of turbulence of incompressible fluid (e.g. Pina 1998). Therefore, it would be valuable to further study the DWT $E/B$ decomposition with the divergence-free vector wavelets." }, "0910/0910.1598_arXiv.txt": { "abstract": "Understanding the details of how the red sequence is built is a key question in galaxy evolution. What are the relative roles of gas-rich vs. dry mergers, major vs. minor mergers or galaxy mergers vs. gas accretion? In a recent paper \\citep{Wild:2009p2609}, we compare hydrodynamic simulations with observations to show how gas-rich major mergers result in galaxies with strong post-starburst spectral features, a population of galaxies easily identified in the real Universe using optical spectra. Using spectra from the VVDS deep survey with $=0.7$, and a principal component analysis technique to provide indices with high enough SNR, we find that 40\\% of the mass flux onto the red-sequence could enter through a strong post-starburst phase, and thus through gas-rich major mergers. The deeper samples provided by next generation galaxy redshift surveys will allow us to observe the primary physical processes responsible for the shut-down in starformation and build-up of the red sequence. ", "introduction": "Recent observations have revealed that since a redshift of around unity the total mass of stars living in red sequence galaxies has increased by a factor of two \\citep[e.g.][]{2004ApJ...608..752B}. At the same time, the stellar mass density of the blue sequence has remained almost constant. The interpretation is that some blue galaxies migrate onto the red sequence after the quenching of their star formation, whilst the remainder continue to form new stars \\citep[e.g.][]{2007ApJ...665..265F}. \\citet{2007A&A...476..137A} measured the net mass flux which has taken place from the blue sequence to the red sequence. This amounts to $9.8\\times10^{-3}$M$_\\odot$/yr/Mpc$^3$, or about $1.4\\times10^4$\\,M$_\\odot$/yr in the VIMOS VLT DEEP Survey (VVDS) volume. Recent star formation history can be used as a tool to identify galaxies as they enter the red sequence. For typical observable times of post-starburst features in VVDS optical galaxy spectra of $\\sim0.35-0.6$\\,Gyr, this could comprise, for example, a few tens of galaxies of stellar mass \\logm$=10.5$ in the VVDS survey. In these proceedings I will summarise the work presented in detail in \\citet{Wild:2009p2609} and ask whether observations of post-starburst galaxies are consistent with the growth of the red-sequence being caused by gas-rich major mergers. ", "conclusions": "" }, "0910/0910.2078_arXiv.txt": { "abstract": " ", "introduction": "\\label{sec:intro} PHL~932 was first noted as a 12th-magnitude blue star during a survey by Haro \\& Luyten (1962). It was studied in more detail by Arp \\& Scargle (1967), who discovered a faint emission nebula surrounding the star which they assumed to be a planetary nebula (PN). The nebula is asymmetrically placed around the ionizing star (Figure~\\ref{fig:neb}), which has a position of $\\alpha,\\delta$ = 00$^{h}$59$^{m}$56.67$^{s}$, +15\\arcdeg 44$^{m}$ 13.7$^{s}$ (J2000). A spectrum of the nebula was obtained by Arp \\& Scargle (1967), who noted strong [O\\,\\textsc{ii}] $\\lambda$3727 and weak H$\\alpha$ emission. The almost complete absence of [O\\,\\textsc{iii}] emission (see \\S\\ref{sec:em_nebula} below) suggests the nebula has a very low ionization parameter, assuming it is photoionized. An investigation of the stellar spectrum by M\\'endez et al. (1988) placed the central star squarely on the extreme horizontal branch (EHB) and allowed the classification of PHL\\,932 as a hot subdwarf~B (sdB) star.\\footnote{M\\'endez et al. (1988) classified the star as ``late sdO'' (or sdOB) based on their detection of HeII $\\lambda$4686 in absorption.} Subdwarf B stars are unlike the typical central stars of PN. They do not evolve to the asymptotic giant branch (AGB), but rather move directly on to the white dwarf cooling curve. Their progenitors are solar-like stars that have somehow shed nearly all of their hydrogen envelopes prior to the onset of core helium fusion (e.g. Heber 1986). Exactly how this occurs is uncertain, although it is possible that some kind of binary interaction is involved (Han et al. 2003). Indeed, De Marco et al. (2004) and Af\\c{s}ar \\& Bond (2005) have claimed the radial velocity of PHL\\,932 is variable, which may indicate binarity (however, see Section~\\ref{sec:binary}). More recently, De Marco (2009) has strongly argued for the general importance of binary interactions for the formation and shaping of PN. Despite the unusual nature of this system, the nebula around PHL~932 has not been studied in much detail since its discovery. In this paper, we present an in-depth analysis of the nebula and the ionizing star based on new and archival multi-wavelength observations. We use the results of this investigation to examine the evolutionary status of the star and to reclassify the nebula. \\begin{figure}[h] \\begin{center} \\includegraphics[width=7cm]{PHL932BRcomb.ps} \\caption{The asymmetric emission nebula around PHL\\,932. The star is at the centre of this combined SuperCOSMOS broadband red and blue image, which is 10\\arcmin\\ across.}\\label{fig:neb} \\end{center} \\end{figure} ", "conclusions": "\\label{sec:disc} We have used a combination of new and archival multi-wavelength observations to show that (a) PHL\\,932 is unlikely to be in a close binary system as previously suggested, and (b) the emission nebulosity surrounding PHL\\,932 is \\emph{not} a PN as has been commonly assumed, nor is it physically associated with the star. We conclude instead that PHL~932 is ionizing a ``clumpy'' region of the ISM as it travels through it; in other words, the emission nebula is simply a Str\\\"omgren zone (or HII region), rather than a PN. We note that another putative PN, EGB~5 (Ellis, Grayson \\& Bond 1984) is also associated with an sdB star. An unpublished MSSSO 2.3-m long-slit spectrum of the nebula taken by one of us (D.J.F) shows a nebula of low excitation, similar to that around PHL~932. Like PHL~932, the morphology of EGB 5 on DSS images militates against it being a bona fide PN, though deep narrowband \\ha\\ images are required for a definitive conclusion. We likewise consider the nebula around EGB~5 to be another candidate Str\\\"omgren zone in the ISM. Such nebulae are seen around hot subdwarfs and white dwarfs when the ISM is sufficiently dense to give an emission measure that facilitates optical detection (cf. Tat \\& Terzian 1999). In future papers in this series, we will show that several other nearby emission nebulae currently assumed to be PN are also Str\\\"omgren zones in the ISM; see Frew \\& Parker (2006) and Madsen et al. (2006) for preliminary results." }, "0910/0910.5372_arXiv.txt": { "abstract": "We present an analysis of the evolution of 8625 Luminous Red Galaxies between $z=0.4$ and $z=0.8$ in the 2dF and SDSS LRG and QSO (2SLAQ) survey. The LRGs are split into redshift bins and the evolution of both the luminosity and stellar mass function with redshift is considered and compared to the assumptions of a passive evolution scenario. We draw attention to several sources of systematic error that could bias the evolutionary predictions made in this paper. While the inferred evolution is found to be relatively unaffected by the exact choice of spectral evolution model used to compute K+e corrections, we conclude that photometric errors could be a source of significant bias in colour-selected samples such as this, in particular when using parametric maximum likelihood based estimators. We find that the evolution of the most massive LRGs is consistent with the assumptions of passive evolution and that the stellar mass assembly of the LRGs is largely complete by $z\\sim0.8$. Our findings suggest that massive galaxies with stellar masses above $10^{11}$M$_\\odot$ must have undergone merging and star formation processes at a very early stage ($z\\gtrsim 1$). This supports the emerging picture of \\textit{downsizing} in both the star formation as well as the mass assembly of early type galaxies. Given that our spectroscopic sample covers an unprecedentedly large volume and probes the most massive end of the galaxy mass function, we find that these observational results present a significant challenge for many current models of galaxy formation. ", "introduction": "Luminous Red Galaxies (LRGs) are arguably some of the brightest galaxies in our Universe allowing us to study their evolution out to much higher redshifts than is possible with samples of more typical galaxies. LRGs are known to form a spectroscopically homogenous population that can be reliably identified photometrically by means of simple colour selections \\citep{Eisenstein:01,Cannon:2SLAQ}. The presence of a strong 4000 \\AA break in these galaxies also means that accurate photometric redshifts can be derived for large samples where spectroscopy is difficult to obtain \\citep{Padmanabhan:05, Collister:MegaZ, Abdalla:MegaZ}. Consequently, samples of luminous red galaxies have emerged as an important data set both for studies of cosmology \\citep{Blake:LRG07,Blake:LRG08,Cabre:LRG09} and galaxy formation and evolution \\citep{Wake:06,Cool:08}. The study of massive galaxies in our Universe is particularly interesting since these galaxies present a long-standing problem for models of galaxy formation. While in the standard $\\Lambda$CDM cosmology, massive galaxies are thought to be built up through successive mergers of smaller systems \\citep{WhiteRees:78}, many observations now suggest that these galaxies were already well assembled at high redshifts \\citep{Glazebrook:04, Cimatti:06, Scarlata:07, Ferreras:09, Pozzetti:09}. At the same time, star formation and merging activity has also been seen in such systems between $z\\sim1$ and the present day \\citep{Lin:04, Stanford:04} and the studies of \\citet{Bell:04} and \\citet{Faber:07} have found evidence that the luminosity density of massive red galaxies has remained roughly constant since z$\\sim$1 implying a build up in the number density of these objects through mergers. Most observational results now support downsizing in star formation - i.e most massive galaxies having lower specific star formation rates than their less massive counterparts - and many galaxy formation models are able to reproduce these observations e.g. by invoking quenching mechanisms such as AGN feedback \\citep{Croton:06, DeLucia:06}. However, the more recently observed trend of downsizing in the mass assembly \\citep{Cimatti:06, Pozzetti:09} presents more of a challenge for galaxy formation models which typically predict that the most massive galaxies were assembled later than their less massive counterparts. Many of the contradictions in observational studies of massive galaxy formation arise principally for two reasons. Firstly, the way in which early-type galaxies are selected in different surveys can be considerably different. Morphological selection compared to colour selection of such objects will almost certainly result in different samples being chosen, particularly at high redshifts. The colour selection is also usually different for different samples of massive galaxies as we will point out later in this paper. Secondly, many of the studies of massive galaxy formation mentioned so far are deep and narrow spectroscopic samples that will suffer from biases due to the effects of cosmic variance. In addition, one has to be careful in interpreting observational results as in current galaxy formation models, star formation and mass assembly in galaxies are not necessarily concomitant. So while the stars in early-type galaxies may have formed very early in the Universe's history, they may have formed in relatively small units and only merged at lower redshifts to create the massive ETGs we see today. Studies of the evolution of luminous red galaxies such as those analysed in this paper have already supported the evidence that massive galaxies have been passively fading and show very little recent star formation \\citep{Wake:06, Cool:08}. In this paper, we extend this work by also considering the mass assembly of these systems. The 2dF and SDSS LRG and QSO (2SLAQ) survey presents a comprehensive improvement in volume for massive galaxy samples. The survey covers an area of 186 deg$^2$ and reaches out to a redshift of 0.8 making it a promising data set to study the evolution of massive red galaxies. Furthermore, much of the 2SLAQ area is now covered by the UKIDSS Large Area Survey (LAS) that provides complementary data in the near infra-red bands for the optically selected LRGs. The advantage of near infra-red data is that the mass-to-light ratios and k-corrections are largely insensitive to the galaxy or stellar type and therefore the total infra-red flux in for example the K-band provides a good estimate of the total stellar mass of the galaxies. This stellar mass estimate allows us to study not only the star formation history but also the mass assembly history of these systems. As is the case for optically selected spectroscopic samples of massive galaxies, the K-band selected surveys have so far been restricted to relatively small and deep patches of the sky \\citep{Mignoli:K20, Conselice:07}. Clearly a large spectroscopic survey of massive galaxies with optical and near infra-red photometry, will allow for better constraints on the evolution of these systems. In this paper, we utilise a spectroscopic sample of colour-selected massive red galaxies from the 2SLAQ survey between redshifts 0.4 and 0.8. Optical photometry is obtained from the Sloan Digital Sky Survey supplemented with near infra-red data from the UKIDSS Large Area Survey (LAS). We consider the evolution of these galaxies in terms of their observed colours, luminosities and comoving number densities focussing particularly on the effects of colour selection on the inferred evolution. \\citet{Wake:06} have presented a comprehensive analysis of the luminosity function of Luminous Red Galaxies using data from both the Sloan Digital Sky Survey and the 2SLAQ Survey. These authors use a subset of data from the 2SLAQ survey in the redshift range 0.5 to 0.6 and by comparing this galaxy population to a lower redshift population at $0.17 3 \\times 10^{11} M_\\odot$ also shows little evidence for evolution between redshifts 0.4 and 0.8 suggesting that these most massive systems were already well assembled at redshifts of 0.8. This is consistent with the emerging picture of downsizing in the mass assembly of massive galaxies.} \\item{The stellar mass function estimate is also relatively insensitive to the choice of spectral evolution model assumed in the calculation of $V_{max}$. However, different choices of the stellar initial mass function will shift the mass function estimate as found in previous studies.} \\item{The comoving number density of LRGs with $M> 3 \\times 10^{11} M_\\odot$ has changed little between redshifts 0.8 and 0.4. The same is true for LRGs with $M > 10^{12} M_\\odot$. This is consistent with other observational results for massive galaxy samples and does not agree with the predictions of most current galaxy formation models. We find that the models of \\citet{Granato:04} may be promising in matching our observations although they are yet to be compared with massive galaxy samples at intermediate redshifts.} \\end{itemize} Overall, our results support the emerging picture of downsizing in both the star formation and mass assembly of early type galaxies and present a significant challenge for current models of galaxy formation. We have shown our findings to be robust to changes in spectral evolution models, cosmic variance and photometric errors and conclude that these new observations of the most massive and most luminous galaxies in our Universe will need to be reconciled with the models if progress is to be made in the field of massive galaxy formation and evolution." }, "0910/0910.2990_arXiv.txt": { "abstract": "We present new near-IR \\hh, CO J=2-1, and CO J = 3-2 observations to study outflows in the massive star forming region IRAS 05358+3543. The Canada-France-Hawaii Telescope \\htwo\\ images and James Clerk Maxwell Telescope CO data cubes of the IRAS 05358 region reveal several new outflows, most of which emerge from the dense cluster of sub-mm cores associated with the \\necluster\\ cluster to the northeast of \\region. We used Apache Point Observatory (APO) JHK spectra to determine line of sight velocities of the outflowing material. Analysis of archival VLA cm continuum data and previously published VLBI observations reveal a massive star binary as a probable source of one or two of the outflows. We have identified probable sources for 6 outflows and candidate counterflows for 7 out of a total of 11 seen to be originating from the \\region\\ clusters. We classify the clumps within \\necluster\\ as an early protocluster and \\swcluster\\ as a young cluster, and conclude that the outflow energy injection rate approximately matches the turbulent decay rate in \\necluster. ", "introduction": "Collimated, bipolar outflows accompany the birth of young stars from the earliest stages of star formation to the end of their accretion phase \\citep[e.g.][]{reipurth2001}. While the birth of isolated low-mass stars is becoming well understood, the formation of massive stars ($>10 \\msun$) and clusters remains a topic of intense study. Observations show that moderate to high-mass stars tend to form in dense clusters \\citep{lada2003}. In a clustered environment, the dynamics of the gas and stars can profoundly impact both accretion and mass-loss processes. Feedback from these massive clusters may play a significant role in momentum injection and turbulence driving in the interstellar medium. Outflows from massive stars are less studied than those from low mass stars largely because massive stars accrete most of their mass while deeply embedded. Therefore, unlike low mass young stars that are accessible in the optical, massive stellar outflows can only be seen at infrared and longer wavelengths. Direct evidence for jets from massive young stellar objects (YSOs) from \\hh\\ or optical emission is generally lacking \\citep[e.g.][]{alvarez2005,kumar2002,wang2003}, although there is evidence that massive stars are the sources of collimated molecular outflows from millimeter observations \\citep[e.g.][]{beuther2002b}. Outflows from massive stars may allow accretion to continue after their radiation pressure would otherwise halt accretion in a spherically symmetric system \\citep{krumholz2009}. They therefore represent a crucial component in understanding how stars above $\\sim$10 \\msun\\ can form. \\region\\ is a double cluster of embedded infrared sources located at a distance of 1.8 kpc in the Auriga molecular cloud complex \\citep{heyer1996} associated with the HII regions Sh-2 231 through 235 at Galactic coordinates around $l,~b$ = 173.48,+2.45 in the Perseus arm. \\necluster\\ is the collection of highly obscured and mm-bright sources slightly northeast of \\swcluster, which is the location of the IRAS 05358+3543 point source and the optical emission nebula (see Figure \\ref{fig:overview_ha}). The IRAS source is probably a blend of the three brightest infrared objects in the MSX A-band and MIPS 24 \\um\\ images, which are located at \\necluster, IR 41, and IR 6. For the purpose of this paper,the whole complex including both sources is referred toas \\region, and otherwise refer to individual objects specifically. Early observations revealed the presence of OH \\citep{Wouterloot1993}, \\HtwoO\\ \\citep{Scalise1989, Henning1992}, and methanol \\citep{Menten1991} masers about an arcminute northeast of the IRAS source, indicating that massive stars are likely present at that location. Near infrared observations revealed the presence of two embedded clusters \\citep{porras2000,jiang2001} labeled \\swcluster\\ for the southwestern cluster associated with the IRAS source, and \\necluster\\ for the northeastern cluster located near the OH, \\HtwoO, and CH$_3$OH masers. Stars identified in \\citet{porras2000} are referred to by the designation ``IR (number)'' corresponding to the catalog number in that paper. \\citet{porras2000} also included scanning Fabry-Perot velocity measurements of the inner $\\sim1$\\arcmin. CO observations revealed broad line wings indicative of a molecular outflow \\citep{casoli1986,shepherd1996}. \\citet{kumar2002} and \\citet{khanzadyan2004} presented narrow band images of 2.12 \\um\\ \\htwo\\ emission that reveled the presence of multiple outflows. Interferometric imaging of CO and SiO confirmed the presence of at least three flows emerging from the northeast cluster centered on the masers \\citep{beuther2002} having a total mass of about 20 \\msun . \\citet{beuther2002} also presented MAMBO 1.2 mm maps and a mass estimate of 610 \\msun\\ for the whole region. \\citet{williams2004} presented SCUBA maps and mass estimates of the clusters of 195/126\\msun\\ for \\necluster\\ and 24/12 \\msun\\ for \\swcluster\\ (850 \\um/450 \\um). \\citet{Zinchenko1997} measured the dense gas properties using the \\ammonia\\ (1,1) and (2,2) lines. They measure a mean density $n \\approx 10^{3.60}$ \\percc, temperature 26.5K, and a mass of 600 \\msun . The total luminosity of the two clusters is about 6300 \\lsun , indicating that the region is giving birth to massive stars \\citep{porras2000}. Millimeter wavelength interferometry with arcsecond angular resolution has revealed a compact cluster of deeply embedded sources centered on the \\HtwoO\\ and methanol maser position \\citep{beuther2002,beuther2007,leurini2007}. \\citet{beuther2002} identified 3 mm continuum cores, labeled mm1-mm3 (shown in Figure \\ref{fig:outflowsh2}). \\citet{beuther2007} resolved these cores into smaller objects. Source mm1a is associated with a cm continuum point source and will be discussed in detail below. \\region\\ has previously been observed at low spatial resolution in the J=2-1 and J=3-2 transitions with the Kosma 3m telescope \\citep{Mao2004}. While the general presence of outflows was recognized and a total mass estimated, the specific outflows were not resolved. \\citet{beuther2002} observed the CO J=6-5, J=2-1, and J=1-0 transitions at moderate resolution in the inner few arcminutes. \\citet{Thomas2008} observed C$^{17}$O in the J=2-1 and J=3-2 transitions with a single pointing using the JCMT. ", "conclusions": "We have presented a multiwavelength study of the \\region\\ star forming region. \\region\\ contains an embedded cluster of massive stars and is surrounded by outflows. The outflows were linked to probable sources and determined that at least one outflow is probably associated with a massive ($\\sim 10 \\msun$) star. Added kinematic information and a wide field view of the infrared outflows has been used to develop a more complete picture of the region. \\begin{itemize} \\item \\necluster\\ is a Protocluster and \\swcluster\\ is a Young Cluster \\item Energy injection on the scales of \\region\\ can maintain turbulence, but on the small scales of the \\necluster\\ protocluster, is inadequate by $\\sim 2$ orders of magnitude. \\necluster\\ is collapsing. \\item there are 11 candidate outflows, 7 of which have candidate counterflows, in the \\region\\ complex \\item there is a probable massive binary with one member of mass 12 \\msun\\ in mm1a, and the other which is the source of Outflow 2 \\item there are at least two moderate-mass ($\\sim$5\\msun) young stars in \\region\\ \\end{itemize} We have identified additional middle- and high-mass young stars with outflows, and presented a case for a high-mass binary system within the millimeter core mm1a." }, "0910/0910.5054_arXiv.txt": { "abstract": "{ Cosmology contributes a good deal to the investigation of variation of fundamental physical constants. High resolution data is available and allows for detailed analysis over cosmological distances and a multitude of methods were developed. The raised demand for precision requires a deep understanding of the limiting errors involved. The achievable accuracy is under debate and current observing proposals max out the capabilities of todays technology. The question for self-consistency in data analysis and effective techniques to handle unknown systematic errors is of increasing importance. An analysis based on independent data sets is put forward and alternative approaches for some of the steps involved are introduced. ", "introduction": "This work is motivated by numerous findings of different groups that partially are in disagreement witch each other. A large part of these discrepancies reflect the different methods of handling systematic errors. Evidently systematics are not yet under control or fully understood. We try to emphasize the importance to take these errors, namely i.e. calibration issues, into account and put forward some measures adapted to the problem. ", "conclusions": "" }, "0910/0910.2397_arXiv.txt": { "abstract": "In this paper bulk viscosity is introduced to describe the effects of cosmic non-perfect fluid on the cosmos evolution and to build the unified dark energy (DE) with (dark) matter models. Also we derive a general relation between the bulk viscosity form and Hubble parameter that can provide a procedure for the viscosity DE model building. Especially, a redshift dependent viscosity parameter $\\zeta\\propto\\lambda_{0}+\\lambda_{1}(1+z)^{n}$ proposed in the previous work by X.H.Meng and X.Dou in 2009\\cite{md} is investigated extensively in this present work. Further more we use the recently released supernova dataset (the Constitution dataset) to constrain the model parameters. In order to differentiate the proposed concrete dark energy models from the well known $\\Lambda$CDM model, statefinder diagnostic method is applied to this bulk viscosity model, as a complementary to the $Om$ parameter diagnostic and the deceleration parameter analysis performed by us before. The DE model evolution behavior and tendency are shown in the plane of the statefinder diagnostic parameter pair \\{$r,s$\\} where the fixed point represents the $\\Lambda$CDM model. The possible singularity property in this bulk viscosity cosmology is also discussed to which we can conclude that in the different parameter regions chosen properly, this concrete viscosity DE model can have various late evolution behaviors and the late time singularity could be avoided. We also calculate the cosmic entropy in the bulk viscosity dark energy frame, and find that the total entropy in the viscosity DE model increases monotonously with respect to the scale factor evolution, thus this monotonous increasing property can indicate an arrow of time in the universe evolution, though the quantum version of the arrow of time is still puzzling. ", "introduction": "Type Ia supernova and other astrophysics observations together indicate that our universe is accelerating now \\cite{sn1}. Different models are proposed to try to describe or understand this surprisingly exotic phenomenon. If we make the assumption that general relativity is still correct to the scale of cosmos, an effective term contributes to negative pressure should be added to the right hand side of Einstein's field equation in the general theory of relativity to explain the recent stage speed-up of our observational universe expansion. This is the basic idea to the so called dark energy concept. So far for the ten years' old DE phenomena there are many both theoretical and observational attempts to understand the mechanism. The introduction of the cosmological constant corresponds to a negative pressure fluid with a specially constant density all the time which has been playing a key role in the universe evolution by uniformly distributed over the whole cosmic space-time and media, but the cosmological constant existence raises serious fundamental physics problem, the so-called cosmological constant problem for both new (coincidence) and old (so tiny) ones. Another class of models try to modify the traditional Einstein's general theory of relativity to the large cosmology scale, by arguing that the recently appearing acceleration phase of the unverse expansion comes from the break down of general relativity in cosmic scale. The so-called $f(R)$ gravity, to mention one for example, which generalizes the Hilbert-Einstein action is categorized in this class \\cite{mw1}. (For more details and references, you may see the recent reviews \\cite{r1} \\cite{r2}). In the context of perfect fluid, some models based on fluid mechanics method are studied extensively, such as the Chaplygin gas and generalized Chaplygin gas models which modify the equation of state, and barotropic fluids dark Energy \\cite{ls}. Also for the purpose to consider more realistic situation in the evolution of the universe, the concept of viscosity is introduced into the investigation of the cosmos evolution from fluid mechanics \\cite{v2} \\cite{v3} \\cite{v7} \\cite{v8} \\cite{v9} \\cite{v10} \\cite{v11}. Earlier attempt \\cite{pc} in this area even ``predicts'' the late acceleration of the universe expansion. The dissipative effect in the fluid is always due to shear and bulk viscosity characterized by shear viscosity parameter $\\eta$ and bulk viscosity parameter $\\zeta$. In the study of cosmology, the shear viscosity disappears in the Friedmann-Robertson-Walker's isotropic and homogeneous framework with the largest spherical symmetry. Only bulk viscosity could play a role in the realistic models. Its effects can be shown from an added correction term with the minus sign to cosmic pressure $\\tilde p=p-3\\zeta H$. This composite formula motives the trial on the connection between bulk viscosity of the cosmic fluid and dark components (energy and matter). As we understand so far that purely gravitational probes can only provide information on a single effective matter-energy fluid, which may consist of the dark energy and (dark) matter composites as well as the least dominated radiation energy at present universe, so for the dominated two functioning dark components we can describe them as a single dark fluid \\cite{Ren}. In this paper, we discuss a concrete unified model building by paying attention on the modification of the cosmic media from the simple assumption as a perfect fluid to a piratical viscous fluid encoded in the energy-stress tensor contents, that is, at the right hand side of the Einstein's field equation. For the bulk viscosity DE model building a remark is needed here. Considering the observational dark energy ingredient fraction today, it takes about $\\frac{2}{3}$ of the whole matter-energy density, so the corresponding pressure provided by the bulk viscosity correction dominantly surpasses the remaining pressure contributions from other cosmic energy-matters. This is obviously different from the traditional non-equilibrium thermodynamics, in which the viscosity contribution is only a little correction to the pressure term. Some researches try to find a mechanism to support such a fluid behavior by a kind of non-standard interaction introduced between the dark matter and energy components \\cite{inter1} \\cite{inter2}. It is certainly important for these attempts to find more support or hints behind cosmic observations in the study of bulk viscosity cosmology along this possible line. For concrete model building, detailed form of viscosity parameter is needed to settle down the theoretical framework. In the reference \\cite{Ren}, a scale factor dependent viscosity is proposed, which is different from the only density dependent form \\begin{equation} \\zeta=\\zeta_{0}+\\zeta_{1}\\frac{\\dot a}{a}+\\zeta_{2}\\frac{\\ddot a}{\\dot a} \\end{equation} It could be shown that it is equivalent to a modified equation of state(EOS) \\begin{equation} p=(\\gamma-1)\\rho+p_{0}+w_{H}H+w_{H2}H^{2}+w_{dH}\\dot H. \\end{equation} And other interesting physical properties have been investigated in that kind of models \\cite{rm} \\cite{con}. In this present paper, by largely extending its contents we will continue the study of a redshift dependent viscosity form $\\zeta\\propto\\lambda_{0}+\\lambda_{1}(1+z)^{n}$ as proposed in the previous work \\cite{md}. We concentrate on the late time singularity discussion and its entropy expression of the bulk viscosity DE model. For the phantom dark energy model, there exists a cosmic singularity in the future cosmos evolution, that is, the so-called cosmic ``doomsday''. As shown below, we could see that this viscosity DE model represents different evolution behaviors, and the future cosmic singularity will disappear under some proper parameter ranges selection. From this view of point, the model has the flexibility to produce either quintessence or phantom properties, that is, we can easily achieve its EoS either larger or less than characteristic -1. An important tool for investigating dark energy model characters nowadays is by the introduction of some geometry quantities, for instance the statefinder diagnostic parameters \\cite{sf}, which are quantities dependent on high order derivatives of the scale factor, such as to the $\\dddot a$. The usual statefinder parameter pair \\{$r,s$\\} of the model concerned is calculated explicitly to demonstrate the viscosity DE model behaviors. In our previous work, deceleration parameter and $Om$ diagnostic parameter \\cite{om} are performed to show the properties. We have found that in statefinder pair plane, our model and $\\Lambda$CDM model could be easily discriminated in some redshift ranges. We prospect that increasing the quality and quantity of measurement data will enhance our ability to make accurate discrimination of different current cosmology models and rule out some. At the same time, new diagnostic methods merit further investigations, especially diagnostic quantity in the higher order perturbation level. The paper is organized as follows: in the second section, some general features and remarks about viscosity dark energy model are given, and we further discuss the redshift dependent model. In the third section, we give the singularity discussion of our model. Some solutions are given there. In Sec. \\textbf{IV}, we calculate entropy of this viscosity model. Finally, conclusion and discussions are presented. We leave the data fitting procedure in the appendix. ", "conclusions": "In this letter we continue and largely extended our previous work on the single and unified bulk viscous fluid as a potential dark energy candidate by presenting an explicit viscosity form to mimic dark energy behaviors and confront it with current observational data sets. The specific feature here is a variable coefficient for the new bulk viscosity form proposed, characterized by two free parameters that can be best fitted by astrophysics observational data sets. the best fitting results have shown that this concrete model could yield theoretical prediction values in an acceptable level by working out the numerical processing to the latest released joint observational data sets. Furthermore, we have performed the statefinder diagnostic parameter analysis to this unified viscosity DE model, finding that in most evolution stages of the unverse, statefinder parameters could be used to obviously distinguish this viscosity DE model from the $\\Lambda$CDM model. But as shown in the figure of the statefinder pair parameter \\{$r,s$\\}, the viscosity DE model evolution trajectory passes the special point corresponding to the well known $\\Lambda$CDM model, where the models degeneration emerges. The new $Om$ parameter diagnostics made previously could be a more powerful tool then, which might discriminate concrete models from the $\\Lambda$CDM model in the whole evolution history. In this present work, we particularly concentrate on the singularity behavior of the unified viscosity dark energy model. We find that different parameter range selection, especially the region of power parameter $n$, influences the finite future singularity. For the $n=0$ case, which corresponds to the case with a constant viscosity, there is an exact solution for the evolution scale factor. For the early universe and the $n<0$ case, the solution of the scale factor $a(t)$ has possessed the same structure. A further study of viscosity effects on the early universe evolution, especially during inflation stage, seems very interesting with rich possibilities. In the context of the unified viscosity DE cosmology, we also calculate the entropy for the total evolution universe, which has been expressed in a complex form with this non-perfect viscosity media. Calculations of the total entropy for the viscosity universe evolution represent that the general second law of thermodynamics holds in the whole or global universe description. The worked out results show that in this complicated context, the total entropy is increasing with the redshift decreasing, or cosmic time flying, including the cosmic expansion accelerating stage as observational data indicating now, which has been plotted in figure 5 with three free parameter values chosen. The monotonous increasing property of the total entropy with cosmic time flowing, the viscosity DE universe evolution may provide us an arrow of time to describe the complex universe changing direction. Though a definite concept for the arrow of time is debating, now very puzzling in different situations or versions: classical physics, quantum physics, cosmology and quantum gravity, especially it is essential for a correct quantum gravity theory to be expected to appear, the thermodynamics second law is generally believed to hold to describe the global evolution of our observational universe. So the reasonable entropy expression is richly encoded helpful information for the concerned system. It is certainly interesting and worthy of further efforts. Dark energy physics involves many fundamental concepts and beyond in our already established \"standard models\" for both particle physics and gravity. Viscosity media seems to relate the matter sector to the geometric gravity side via the Einstein's general relativity equation. Actually it also can be reconstructed effectively from the left side to the right side of the equation by modified gravity or extra dimension models for example, too, which is also intriguing. With the available and upcoming high quality and increasingly large amount of astrophysics observational data, especially the good low redshift SNe Ia data sets with less uncertainty for possible errors from the dust effects alleviated under control we expect the ten-years old mysterious dark energy problem will be pinned down not too long compared with the long time standing unsolved dark matter mystery. Maybe the practically unified viscosity DE model or the like can provide an economic mechanism to answer the both uncovered secrets in one dark sector, a tale for the two mysteries." }, "0910/0910.3672_arXiv.txt": { "abstract": "{We present results of variability search in the field of the young open cluster NGC\\,1502. Eight variable stars were discovered. Of six other stars in the observed field that were suspected for variability, we confirm variability of two, including one $\\beta$~Cep star, NGC\\,1502-26. The remaining four suspects were found to be constant in our photometry. In addition, $UBVI_{\\rm C}$ photometry of the well-known massive eclipsing binary SZ~Cam was obtained. The new variable stars include: two eclipsing binaries of which one is a relatively bright detached system with an EA-type light curve, an $\\alpha^2$\\,CVn-type variable, an SPB candidate, a field RR Lyr star and three other variables showing variability of unknown origin. The variability of two of them is probably related to their emission in H$\\alpha$, which has been measured by means of the $\\alpha$ index obtained for 57 stars brighter than $V \\approx$16 mag in the central part of the observed field. Four other non-variable stars with emission in H$\\alpha$ were also found. Additionally, we provide $VI_{\\rm C}$ photometry for stars down to $V =$ 17~mag and $UB$ photometry for about 50 brightest stars in the observed field. We also show that the 10-Myr isochrone fits very well the observed color-magnitude diagram if a distance of 1~kpc and mean reddening, $E(V-I_{\\rm C}) =$ 0.9~mag, are adopted.} {stars: early-type -- stars: emission-line, Be --- open clusters and associations: individual: NGC\\,1502} ", "introduction": "This is the seventh paper in the series in which we present results of the variability search for B-type variables (mainly $\\beta$~Cep, SPB and massive binaries) in the young open clusters of the Northern Hemisphere. The results for previously studied clusters were briefly summarized by Pigulski \\etal (2002). In addition to the discovery or verification of variability of about 70 stars, mostly members of the investigated clusters, it appeared that the incidence of $\\beta$~Cep stars in the northern clusters is lower than in the southern ones. This was interpreted in terms of the metallicity gradient in the Galaxy. One of the main motivations for the search is also a selection of clusters containing large numbers of pulsators suitable for asteroseismology. An example is the previous paper of the series (Ko\\l{}aczkowski \\etal 2004), in which we investigated the open cluster NGC\\,6910 discovering four $\\beta$~Cep stars. This prompted us to choose NGC\\,6910 as a target of a two-year photometric and spectroscopic campaign. The data from the campaign are in the course of reduction, but preliminary analysis already increased the number of $\\beta$~Cep stars in this cluster to seven (Pigulski \\etal 2007) making it an excellent target for seismic modelling and one of very few which are rich in $\\beta$~Cep stars. In the present paper, we report the results of the variability search in the field of another young open cluster, NGC\\,1502. For the first time, the data from the new CCD camera, with a much larger field of view, are included. ", "conclusions": "The open cluster NGC\\,1502 contains only a single low-amplitude $\\beta$~Cep star (NGC\\,1502-26). This appears to be in a contrast to NGC\\,6910 (Ko{\\l}aczkowski \\etal 2004, Pigulski \\etal 2007) which is not very rich in stars either but harbors as many as seven $\\beta$~Cep stars. Whether the difference can be interpreted in terms of a difference in metallicity is not known. Unfortunately, the only metallicity determination that can be related to NGC\\,1502 is that from Str\\\"omgren photometry by Eggen (1985) for HD\\,25056 in Cam~OB1. He reported [Fe/H] $=$ 0.16~dex for this star. The value of [Fe/H] is rather uncertain. In addition, the star might not be related to NGC\\,1502 at all. Definitely, a spectroscopic determination of abundances of stars in the cluster are highly desirable. SZ Cam remains one of the most interesting stars of the cluster being an excellent case of a massive hierarchical system that are found frequently in the cores of open clusters. Thorough studies of such systems may shed light on the role of binaries in massive star formation in general. SZ~Cam was also succesfully used to derive the distance to the cluster (Gorda \\etal 2007) and we may hope that further observations will improve this determination. The relatively bright EA-type system (NGC\\,1502-7) we discovered might be also used for this purpose. \\Acknow{This research has made use of the WEBDA database, operated at the Institute for Astronomy of the University of Vienna. This work was supported by two grants: N\\,N203\\,302635 from Polish MNiSzW and Chilean Proyecto FONDECYT Nr 3085010. We thank Prof.~M.\\,Jerzykiewicz for his comments made upon reading the manuscript. We also thank J.\\,Molenda-\\.Zakowicz, G.\\,Kopacki and Z.\\,Ko{\\l}aczkowski for making some observations of the cluster.}" }, "0910/0910.4863_arXiv.txt": { "abstract": "The centers of most galaxies in the local universe are occupied by compact, barely resolved sources. Based on their structural properties, position in the fundamental plane, and integrated spectra, these sources clearly have a stellar origin. They are therefore called \"nuclear star clusters\" (NCs) or \"stellar nuclei\". NCs are found in galaxies of all Hubble types, suggesting that their formation is intricately linked to galaxy evolution. Here, I review some recent studies of NCs, describe ideas for their formation and subsequent growth, and touch on their possible evolutionary connection with both supermassive black holes and globular clusters. ", "introduction": "The nuclei of galaxies are bound to provide ``special'' physical conditions because they are located at the bottom of the potential well of their host galaxies. This unique location manifests itself in various distinctive phenomena such as super-massive black holes (SMBHs), active galactic nuclei (AGN), central starbursts, or extreme stellar densities. The evolution of galactic nuclei is closely linked to that of their host galaxies, as inferred from a number of global-to-nucleus scaling relations discovered in the last decade. Recently, observational and theoretical interest has been refocused onto the compact and massive star clusters found in the nuclei of galaxies of all Hubble types. These ``nuclear star clusters'' (NCs) are intriguing objects that are linked to a number of research areas: i) they are a promising environment for the formation of massive black holes because of their extreme stellar density, ii) they may also constitute the progenitors of at least some halo globular clusters via ``NC capture'' following the tidal disruption of a satellite galaxy, and iii) their formation process is influenced by (and important for) the central potential, which in turn governs the secular evolution of their host galaxies. In what follows, I briefly summarize what has been learned about NCs over the last few years, describe some proposed formation mechanisms of NCs, and discuss the new paradigm of ``central massive objects'' which links NCs with SMBHs in galactic nuclei. Lastly, I briefly mention a scenario in which NCs may be the progenitors of (some) globular clusters. ", "conclusions": "" }, "0910/0910.3444_arXiv.txt": { "abstract": "Conventionally, long GRBs are thought to be caused by the core collapses of massive stars. During the lifetime of a massive star, a stellar wind bubble environment should be produced. Furthermore, the microphysics shock parameters may vary along with the evolution of the fireball. Here we investigate the variation of the microphysics shock parameters under the condition of wind bubble environment, \\textbf{and allow the microphysics shock parameters to be discontinuous at shocks in the ambient medium. It is found that our model can acceptably reproduce the rebrightenings observed in GRB afterglows, at least in some cases.} The effects of various model parameters on rebrightenings are investigated. The rebrightenings observed in both the R-band and X-ray afterglow light curves of GRB 060206, GRB 070311 and GRB 071010A are reproduced in this model. ", "introduction": "Gamma-ray bursts (GRBs) are attractive astrophysics phenomena and had puzzled astronomers for about twenty-four years after their discovery in 1973 (Klebesadel et al. 1973). The discovery of long-lived, multi-band counterparts of GRBs, namely, the afterglows of GRBs, in 1997 is a watershed in GRB research (Costa et al. 1997; Van Paradijs et al. 1997; Frail et al. 1997). Soon after that, the so-called fireball model is recognized as the standard model in view of the fact that it can explain most features of GRB observations well. However, as the advance of observation techniques and the accumulation of observational data, especially after the launch of \\emph{Swift} satellite (Gehrels 2004), a lot of unexpected behaviours appear in GRB afterglows, such as the canonical steep-shallow-normal decay and flares in X-ray afterglows, the flattish decay phase and various rebrightenings in optical afterglows (For review, see Zhang 2007). In fact, there are more and more rebrightenings detected in GRB afterglows, including GRB 970508 (Galama et al. 1998a), GRB 990123 (Sari \\& Piran 1999), GRB 021004 (Lazzati et al. 2002), GRB 030329 (Berger et al. 2003), GRB 050525A (Blustin et al. 2006), GRB 050820A (Cenko et al. 2006), GRB 050721 (Antonelli et al. 2006), GRB 060206 (Stanek et al. 2007; Liu et al. 2008), GRB 070125 (Updike et al. 2008), GRB 070311 (Guidorzi et al. 2007a), GRB 071003 (Perley et al. 2008), GRB 071010A (Covino et al. 2008). Many different mechanisms have been proposed to explain these rebrightenings, such as density jump (Lazzati et al. 2002; Dai \\& Wu 2003; Tam et al. 2005), energy injection (Huang et al. 2006), two-component jet (Huang et al. 2004, 2006; Liu et al. 2008), reverse shock (Sari \\& Piran 1999), reverberation of the energy input measured by prompt emission (Vestrand et al. 2006), turn-on of the external shock (Stanek et al. 2007; Molinari et al. 2007), large angle emission (Panaitescu \\& Kumar 2007), and spectral peak of existing forward shock (Shao \\& Dai 2005). Among all the mechanisms, the density jump model needs to be paid special attention to, because density jump in surrounding medium of GRBs is a very natural hypothesis. Since there are more and more examples indicating that some GRBs are associated with Type Ic supernovae (e.g. SN 1998bw/GRB980425, Galama et al. 1998b; SN2003dh/GRB030329, Price et al. 2003) and many host galaxies of GRBs are in process of active star formation (Fruchter et al. 1999; Djorgovski et al. 1998), it is believed that the progenitors of long GRBs are massive Wolf-Rayet (WR) stars (Woosley 1993). Massive stars usually produce very strong stellar wind to push the initial interstellar medium (ISM) in their neighborhood away during their lifetime. The surrounding of these GRBs should be a wind bubble following the density profile $\\rho \\propto r^{-2}$ rather than the usual homogeneous ISM. Several authors (Castor et al. 1975; Weaver 1977; Ramirez-Riuz et al. 2001) further found that beyond some typical radius of the wind bubble, the swept-up mass is too large to be pushed by the wind. As a result, the wind materials pile up at the edge of the wind bubble to form a density jump. Lazzati et al. (2002) proposed that a density jump can lead to a rebrightening in the afterglow. In usual case, when the observing frequence is between the peak frequence ($\\nu_{\\rm m}$) and the cooling frequence ($\\nu_{\\rm c}$), the amplitude of the rebrightening should be proportional to the square root of the density contrast. However, if the density contrast is too high, the rebrightening will be weakened, since $\\nu_{\\rm c}$ will decrease and become less than the observing frequence. In the more detailed numerical simulations by Huang et al. (2007), no obvious rebrightenings can be discriminated when the density jump amplitude is set to 100, which suggests that the simple density jump model is not an ideal mechanism to produce the rebrightenings. In the recent studies by Nakar \\& Granot (2007) and van Eerten et al. (2009), a full treatment of the transient features at the jump moment, even including the thickness of the blastwave, was considered. Thus these studies should be a more accurate approximation to the reality. Interestingly, no obvious rebrightenings associated with the density jumps are found. In short, a simple density jump model is difficult to explain the observed rebrightenings in GRB afterglows. On the other hand, the early afterglows of most GRBs exhibit flattish decays with $\\alpha < 0.8$, where $\\alpha$ is defined as $F_{\\nu} \\propto t^{-\\alpha}$. This is difficult to explain using the standard fireball model when the density profile is $\\rho \\propto r^{-2}$. Some previous works have suggested that the microphysics parameters may vary during the evolution of the fireball (Rossi \\& Rees 2003; Ioka et al. 2006; Fan \\& Piran 2006; Panaitescu et al. 2006; Granot, K\\\"{o}nigl \\& Piran 2006). We believe that this kind of variation could resolve these problems. In this paper, we show that the observed rebrightenings in GRB afterglows can be well reproduced by assuming varying microphysics shock parameters associated with the wind bubble environments. The outline of our paper is as follows: in \\S 2 we introduce our model detailedly. We then numerically investigate the effects of various parameters on the afterglows in \\S 3, and reproduce the R-band and X-ray afterglow light curves of GRB 060206, GRB 070311 and GRB 071010A in \\S 4. Our discussion and conclusions are presented in \\S 5. We use an assumptive cosmology of $H_{\\rm 0} = 65$ ${\\rm km}$ ${\\rm s}^{-1}$ ${\\rm Mpc}^{-1}$, $\\Omega_{\\rm M} = 0.30$ and $\\Omega_{\\rm \\Lambda} = 0.70$ throughout the paper. ", "conclusions": "In the standard fireball model, it is assumed that the circum-burst medium density is constant or following a single $\\rho \\propto r^{-2}$ law. The microphysics shock parameters are also usually assumed invariable during the evolution of the fireball. These models can basically explaining many pre-$Swift$ GRB afterglows, whose spectrum and light curves can be approximated as broken power-law functions (Panaitescu \\& Kumar 2001a, 2001b; Yost et al. 2003). The launch of \\emph{Swift} satellite (Gehrels 2004) makes it possible to observe early afterglows of many GRBs in the first few hours after the trigger. Many remarkable and unexpected features are found, such as marked rebrightenings and the flattish decay phase in early GRB afterglows. In fact, the standard fireball model is obviously too simplified. If long GRBs indeed originate from the death of massive stars, the environment of GRBs should have complex structures rather than have a constant or single power law profile. Some analytical (Castor et al. 1975; Weaver 1977; Ramirez-Riuz et al. 2005; Pe'er and Wijers 2006) and numerical (Ramirez-Riuz et al. 2001) studies suggest that the circum-burst density profile should be a wind bubble, associated with a few density jumps. On the other hand, the microphysics parameters, such as $\\xi_{\\rm e}$ and $\\xi_{\\rm B}$, may vary during the evolution of the fireball (Rossi \\& Rees 2003; Ioka et al. 2006; Fan \\& Piran 2006; Panaitescu et al. 2006; Granot, K\\\"{o}nigl \\& Piran 2006). Actually, the fast decrease of the cooling frequency $\\nu_c$ in some GRBs suggests that $\\xi_{\\rm B}$ may be evolving (Panaitescu et al. 2006). Some previous studies also suggest that $\\xi_{\\rm e}$ and $\\xi_{\\rm B}$ may be different for the forward shock and the reverse shock (Fan et al. 2002; Wei, Yan \\& Fan 2006), and may also be different for Region (1) and Region (2) (Gebdre et al. 2007; Kamble, Resmi \\& Misra 2007). So, $\\xi_{\\rm e}$ and $\\xi_{\\rm B}$ may depend on the strength of the shock and the environment. In our study, we combine the wind bubble environments and the change of microphysics shock parameters together. Comparing with the standard fireball model, our model has three more parameters (i.e. $r_{\\rm wind}$, $\\alpha1$ and $\\alpha2$). We find that this model can produce the observed rebrightenings and flattish decay in GRB afterglows successfully. We have shown that the observed rebrightenings in both the optical and X-ray afterglow light curves of GRB 060206, GRB 070311 and GRB 071010A can be well explained by our model. We have also investigated the effects of various parameters on the light curves numerically. We can imagine that if the stellar wind produced by the progenitor is very strong, or the launching speed of the stellar wind $v_{\\rm w}$ is small, or the value of $r_{\\rm wind}$ is large enough, there will be no rebrightening within the usual observation time. So, basically the wind bubble environment can also give birth to a steadily decaying afterglow that shows no rebrightening. In our work, we neglect the effect of the reverse shock. According to the study by Dai \\& Lu (2002), when an ultrarelativistic blast wave interacts with a density jump medium, the resulting reverse shock is relativistic only if the amplitude of the density jump is much larger than 21. In our model, the amplitude of the density jump is only 4, so the corresponding reverse shock is Newtonian. The emission from the Newtonian reverse shock is very weak, and can be omitted. Moreover, although we use an abrupt density jump, the actual increase of density may be gradual. In this situation, the emission from the reverse shock will even be much weaker. Currently, a complete understanding of the microphysical processes in the relativistic shocks is still lacking. Using the derived microphysical parameters from GRB modeling, people may be able to get some constraints on the shock physics. The derived parameter values of $\\dot M$ and $n_{\\rm ISM}$ are also very important. They are closely related to the evolution and the environment of the massive star. They may give some hints on the characters of the progenitor, such as the initial main-sequence mass and the metallicity. So, from the modeling parameters, we can also obtain some useful information about GRB origin." }, "0910/0910.4578_arXiv.txt": { "abstract": "The Cosmic Neutrino Background ({\\cnub}) anisotropy is calculated for massive neutrino states by solving the full Boltzmann equation. The effect of weak gravitational lensing, including the Limber approximation, is also derived for massive particles, and subsequently applied to the case of massive neutrinos. ", "introduction": "The Cosmic photon Microwave Background (CMB) is currently our main source of information about the physical content of the Universe. Observations of the CMB anisotropy provides detailed information about the curvature of the Universe, the matter content, and a plethora of other parameters \\cite{Komatsu:2008hk}. Standard model physics likewise predicts the presence of a Cosmic Neutrino Background (\\cnub) with a well defined temperature of $T_\\nu \\sim (4/11)^{1/3} T_\\gamma$. While it remains undetected in direct experiments, the presence of the {\\cnub} is strongly hinted at in CMB data. The homogeneous C$\\nu$B component has been detected at the 4-5$\\sigma$ level in the WMAP data (see e.g.\\ \\cite{Komatsu:2008hk,Hamann:2007pi,de Bernardis:2007bu,Ichikawa:2008pz,Hamann:2008we,Popa:2008tb}). Furthermore, this component is known to be free-streaming, i.e. to have an anisotropic stress component consistent with what is expected from standard model neutrinos (see \\cite{Bashinsky:2003tk,Trotta:2004ty,Bell:2005dr,DeBernardis:2008ys,Basboll:2008fx,Hannestad:2004qu,Friedland:2007vv}). Finally the standard model neutrino decoupling history is also confirmed by Big Bang Nucleosynthesis (BBN), the outcome of which depends on both the energy density and flavour composition of the {\\cnub}. While this indirect evidence for the presence of a {\\cnub} is important, a direct detection remains an intriguing, but almost impossible goal. The most credible proposed method is to look for a peak in beta decay spectra related to neutrino absorption from the {\\cnub} \\cite{Weinberg:1962zz,Cocco:2007za,Blennow:2008fh}, although many other possibilities have been discussed \\cite{Weiler:1982qy,Stodolsky:1974aq,Gelmini:2004hg,Ringwald:2004np,Fodor:2002hy,Duda:2001hd,Langacker:1982ih,Cabibbo:1982bb}. The neutrino absorption method was first investigated by Weinberg \\cite{Weinberg:1962zz}, based on the possibility that the primordial neutrino density could be orders of magnitude higher than normally assumed due to the presence of a large chemical potential. Although a large chemical potential has been ruled out because it is in conflict with BBN and CMB \\cite{Pastor:2001iu,Pastor:2008ti,Simha:2008mt,Wong:2002fa,Abazajian:2002qx}, the method may still work and recently there has been renewed interest in detecting the {\\cnub} using beta unstable nuclei. Although the direct detection of the {\\cnub} is already very challenging, one might speculate on the possibility that in the more distant future anisotropies in the {\\cnub} will be detectable. For massless neutrinos the calculation of the {\\cnub} proceeds in a way which is almost identical to the standard CMB calculation. The massless {\\cnub} anisotropy spectrum was first presented in \\cite{Hu:1995fqa} and subsequently calculated in \\cite{Michney:2006mk} in a highly simplified way which contains some, but not all of the essential physics. For massive neutrinos the calculation is much more complicated, and the {\\cnub} anisotropy is changed considerably: If the mass is sufficiently high, neutrino velocities can be as low as the escape velocities of galaxies. In this case the {\\cnub} is entirely determined by non-linear gravitational clustering. The current thermal velocity of a non-relativistic homogeneous neutrino background is given roughly by $\\langle v \\rangle \\sim 1500 \\, {\\rm km} \\, {\\rm s}^{-1} \\left(\\frac{0.1 \\, {\\rm eV}}{m_\\nu}\\right)$, which should be compared to gravitational streaming velocities which are up to $\\sim 1000 \\, {\\rm km} \\, {\\rm s}^{-1}$. For small masses (i.e.\\ non-degenerate, $m_\\nu \\lwig 0.1$ eV) it is possible to make a calculation which is analogous to what is done for the CMB. As will be explained later the {\\cnub} spectrum is shifted to larger angular scales, mainly because the much shorter distance to the neutrino last scattering surface changes the relation between angular scales and length scales. Furthermore the amplitude of the anisotropy greatly increases at small multipoles because the gravitational source term becomes much more important as neutrinos become increasingly non-relativistic at late times. In addition to this change in the primary {\\cnub} spectrum, the effect of gravitational lensing is also very different from the case of massless neutrinos. As is the case for the primary spectrum, gravitational lensing becomes increasingly important at low $l$ as the mass increases. A detailed calculation of the lensing of massive neutrinos is presented in section 3. First, however, we derive the necessary equations for the primary {\\cnub} anisotropy in section 2 and present a numerical calculation of the {\\cnub} power spectra for different masses. In section 4 we combine the results of sections 2 and 3 to derive the form of the lensed massive {\\cnub}. Finally, section 5 contains a discussion of our results and our conclusions. ", "conclusions": "We have calculated the anisotropy of the {\\cnub} in linear theory which applies to neutrino masses of less than $\\sim 0.1$ eV. For massless neutrinos the power spectrum of {\\cnub} fluctuations closely resembles the usual CMB spectrum, but with the baryon-photon acoustic oscillations absent. At high $l$ the neutrino spectra are almost identical, independent of the neutrino mass. The reason is that at high $l$ all neutrinos are dominated by free-streaming which in $l$-space has approximately the same impact for all masses. For smaller $l$-values the anisotropy increases dramatically as the mass increases, because the gravitational source term becomes much more important at late times for massive particles. This initially increases the lowest multipoles but via the Boltzmann hierarchy the effect quickly propagates to higher $l$. We then proceeded to calculate the effect of weak gravitational lensing for massive neutrinos and found it to be much stronger at low $l$, and correspondingly weaker at high $l$, as compared to the massless case. Finally we calculated the effect of lensing on the primary {\\cnub} spectra and found the effect to be unimportant (with relative changes at the per mille level up to $l \\sim 50$), but with some differences depending on the neutrino mass. It is worth mentioning that any direct experimental measurement of the {\\cnub} anisotropy will most likely measure flavour states, not mass states. The actual anisotropy measured will therefore be a superposition of anisotropies for three different mass states, weighed with their individual flavour content. We should finally again stress that our results are only valid for masses of $\\lwig 0.1$ eV. For higher masses linear perturbation theory breaks down because neutrino streaming velocities become comparable to the typical gravitational flow velocities so that a significant fraction of neutrinos are bound in structures. In this case the {\\cnub} spectrum must be found from $N$-body simulations of neutrino clustering \\cite{Brandbyge1}. This can also be seen from the fact that the anisotropy at low $l$ is a factor $\\sim 10^9$ higher for $m_\\nu=0.1$ eV than for massless neutrinos. Since the anisotropy for massless particles corresponds to $\\delta \\rho/\\rho \\sim 10^{-5}$ the corresponding $\\delta \\rho/\\rho$ for 0.1 eV neutrinos is of order one, indicating that perturbation theory breaks down." }, "0910/0910.1731_arXiv.txt": { "abstract": "The mass accretion rate of transonic spherical accretion flow onto compact objects such as black holes is known as the Bondi accretion rate, which is determined only by the density and the temperature of gas at the outer boundary. A rotating accretion flow has angular momentum, which modifies the flow profile from the spherical Bondi flow, and hence its mass accretion rate, but most work on disc accretion has taken the mass flux to be a given with the relation between that parameter and external conditions left uncertain. Within the framework of a slim $\\alpha$ disk, we have constructed global solutions of the rotating, viscous hot accretion flow in the Paczy\\'{n}ski-Wiita potential and determined its mass accretion rate as a function of density, temperature, and angular momentum of gas at the outer boundary. We find that the low angular momentum flow resembles the spherical Bondi flow and its mass accretion rate approaches the Bondi accretion rate for the same density and temperature at the outer boundary. The high angular momentum flow on the other hand is the conventional hot accretion disk with advection, but its mass accretion rate can be significantly smaller than the Bondi accretion rate with the same boundary conditions. We also find that solutions exist only within a limited range of dimensionless mass accretion rate $\\dot{m} \\equiv \\dot{M}/\\dot{M}_B$, where $\\dot{M}$ is the mass accretion rate and $\\dot{M}_B$ the Bondi accretion rate: When the temperature at the outer boundary is equal to the virial temperature, solutions exist only for $0.05 \\la \\dot{m} \\le 1$ when $\\alpha=0.01$. We also find that the dimensionless mass accretion rate is roughly independent of the radius of the outer boundary but inversely proportional to the angular momentum at the outer boundary and proportional to the viscosity parameter, $\\dot{m} \\simeq 9.0\\ \\alpha \\lambda^{-1}$ when $0.1 \\la \\dot{m} \\la 1$, where the dimensionless angular momentum measure $\\lambda \\equiv l_{out}/l_B$ is the specific angular momentum of gas at the outer boundary $l_{out}$ in units of $l_B \\equiv GM/c_{s,out}$, $M$ the mass of the central black hole, and $c_{s,out}$ the isothermal sound speed at the outer boundary. ", "introduction": "The amount of mass gravitationally accreted to the compact objects such as black holes is determined by the conditions of gas around the compact objects. The case when gas has a spherically symmetric distribution, a polytropic pressure-density relation, and no angular momentum was solved by Bondi (1952), who found that the mass accretion rate for the transonic accretion is determined only by the density and the temperature of surrounding gas, both assumed to approach constant values far from the accreting objects. This rate, known as the Bondi accretion rate, is widely used as `the mass accretion rate' in a variety of accretion problems. As a function of the density $\\rho_\\infty$ and the isothermal sound speed $c_{s,\\infty}$ of gas at infinity, the Bondi rate for pure hydrogen gas is \\begin{equation}\\label{eq:Bondi_rate} \\dot{M}_B = 4 \\pi \\Lambda \\frac{(GM)^2 \\rho_\\infty}{\\gamma^{3/2}c_{s,\\infty}^3}, \\end{equation} where $G$ is the gravitational constant, $M$ the mass of the central object, and $\\gamma$ the adiabatic index of accreting gas. The constant $\\Lambda(\\gamma)$ is 0.25 for $\\gamma=5/3$ and 1.12 for $\\gamma=1$. If we define the Bondi radius as $r_B \\equiv GM/c_{s,\\infty}^2$, then $\\dot{M}_B = \\Lambda \\gamma^{-3/2} 4\\pi r_B^2 \\rho_\\infty c_{s,\\infty}$. Gravity starts to dominate over gas pressure inside the Bondi radius and the Bondi accretion rate is roughly the mass flux of gas infalling into a sphere of radius $r_B$ with a velocity equal to the sound speed at infinity. However, many astrophysical accretion flows are expected to have certain amount of angular momentum, and the angular momentum will surely affect the flow properties, including the mass accretion rate. Surprisingly, the rate of mass accretion for these rotating, viscous accretion flows for given external boundary conditions, has not been studied yet. We might expect that if the rotational support is negligible at the Bondi radius, the effect of angular momentum to mass accretion rate would be small. But the accretion rate would diminish as the dimensionless measure of rotation $\\lambda \\equiv l_{out}/l_B$ approaches unity, where $l_{out}$ is the specific angular momentum of gas at the outer boundary and $l_B \\equiv r_B c_{s,\\infty} $ is the representative angular momentum expected at the Bondi radius. Although the accretion problem can be reduced to local one in the limit of thin accretion disc where the radial velocity is negligible (Shakura \\& Sunyaev 1973), accretion flow is fundamentally global in the sense that the flow structure is only globally determined. As pointed out by Yuan (1999), the flow property can be qualitatively affected by the outer boundary conditions. The dependence of the mass accretion rate on boundary conditions can only be addressed by constructing global solutions. There have been many studies on the global solutions of rotating accretion flow (e.g., Narayan, Kato, \\& Honma 1997; Chen, Abramowicz, \\& Lasota 1997; Nakamura et al. 1997; Lu, Gu, \\& Yuan 1999) and the effects of outer boundary conditions on accretion flow (Yuan 1999; Yuan et al. 2000), but most have focused on the flow structures and emission properties for a given mass accretion rate, and the determination of how that rate is fixed by external circumstances remains to be addressed. In this work, we focus on the mass accretion rate of the accretion flow, especially its dependence on the density, temperature, and the angular momentum of gas at the outer boundary. We do this by constructing the simplest global, transonic solutions within the slim disk formalism (Abramowicz et al. 1988). Slim disk formalism uses vertical integration to take into account the finite thickness of the accretion disk, and more importantly allows for the radial motion of the accretion flow and the critical point unlike the thin disk approximation (Shakura \\& Sunyaev 1973). ", "conclusions": "\\subsection{Flow properties} Figure 1 shows typical solutions with different amount of angular momentum: high angular momentum (solid lines), intermediate angular momentum (dotted lines), and low angular momentum (dashed lines). By high angular momentum, we mean $\\lambda \\sim 1$, by low angular momentum, $\\lambda \\ll 1$, and by the intermediate angular momentum, between the two limits. The density and the temperature of gas at the outer boundary are the same for the three solutions: $T_{out} = 2.75\\times 10^{9} \\K$ and $\\rho_{out} = 1.5\\times 10^{-6} \\rho_0$ at $r_{out} = 10^3 r_{Sch}$. The dimensionless angular momenta of gas at the boundary and the mass accretion rate are: $\\lambda = 1.7$ and $\\dot{m} = 0.05$ for high angular momentum flow; $\\lambda = 0.27$ and $\\dot{m} = 0.386$ for intermediate angular momentum flow; $\\lambda = 0.14$ and $\\dot{m} = 0.643$ for low angular momentum flow. Most of the thermal energy is advected with the flow rather than radiated away, and temperature stays close to the virial value. The flow with high angular momentum becomes supersonic at small radius $r_{cr} = 2.2 r_{Sch}$ (solid line in Fig. 1a) and the angular momentum (solie line in Fig. 1c) is close to or a few times lower than the Keplerian value (short-long dashed line in Fig. 1c). The radial velocity is significant but still smaller than the near free-fall velocity of spherical Bondi flow. This is a typical hot, radiatively inefficient advection-dominated accretion flow (ADAF) that has been extensively studied (e.g., Narayan \\& Yi 1994, 1995; Abramowicz et al. 1995; Narayan, Kato, \\& Honma 1997; Chen, Abramowicz, \\& Lasota 1997). In spherical accretion, the flow passes the critical point at a much larger radius $r_{cr} \\sim r_B$, and the flow becomes supersonic inside $\\sim r_B$. But in rotating viscous accretion flow, a large part of the flow is subsonic well inside $r_B$ because the rotation of the flow balances out gravity, and the value of $r_{cr}$ depends on the angular momentum of the flow. A low angular momentum flow (dashed line in Fig. 1a) has a critical point at a much larger radius $r_{cr} = 17 r_{Sch}$ compared to the high angular momentum one. It has a larger radial infall velocity at the outer boundary as well because of smaller centrifugal force (Fig. 1b), and the mass accretion rate is higher than that of the high angular momentum flow. These characteristics show that this low angular momentum flow has more resemblance to the spherical flow than to the disk flow as is expected. This type of flow solution is first discovered by Yuan (1999) and its dynamical and thermal properties have been comprehensibly studied by Lu, Gu, \\& Yuan (1999) and Yuan et al. (2000). The flow with intermediate angular momentum (dotted lines) has a critical point at $r_{cr} = 3.2 r_{Sch}$, and shows intermediate characteristics between the high and low angular momentum flow. The flow properties change from disklike to quasi-spherical as the angular momentum of the flow decreases. This is similar to the case of inviscid rotating accretion flow, which becomes disklike or quasi-spherical, depending on the angular momentum (Abramowicz \\& Zurek 1981). This transition of accretion flow from disklike to quasi-spherical in terms of critical radius position is described in detail by Yuan et al. (2000). \\subsection{Mass accretion rate} In spherical accretion, the outer boundary is naturally selected to be outside the critical point, where the density and the temperature of gas stay roughly constant. In rotating viscous flow, the outer boundary is not naturally associated with the critical point which is at much smaller radius. So we consider two choices for the outer boundary radius: one at the Bondi radius, $r_B(T_{out})$, for given $T_{out}$ and the other at a fixed radius regardless of $T_{out}$. The former choice is equivalent to setting the outer boundary at a virial temperature if we define the virial temperature as $T_{vir}(r) \\equiv GMm_p / (2 k r)$. While original Bondi solutions are expressed in terms of the density and temperature of gas at infinity, our solutions start at a finite radius. But since the density and temperature of gas vary little from infinity to the Bondi radius in the Bondi solutions, we compare our solutions for given density and temperature of gas at finite $r_{out}$ against the Bondi solutions with the same density and temperature at infinity. We first discuss the case where the outer boundary is set to be the Bondi radius, $r_{out} = GM/c_{s,out}^2 \\propto T_{out}^{-1}$ for given $T_{out}$, so that the solutions can be meaningfully compared with the Bondi solutions. The density at the outer boundary is either $\\rho_{out} = 1.5 \\times 10^{-6} \\rho_0$ or $1.5 \\times 10^{-9} \\rho_0$. Since the whole set of equations can be rescaled in density except the cooling function $q^-$, the mass accretion rate then is simply proportional to $\\rho_{out}$ when cooling is not important. In such case, the flow profile depends only on $\\dot m$ and not on the specific value of $\\rho_{out}$. However, as the mass accretion rate $\\dot M$ approaches 0.1 times the Eddington mass accretion rate, $\\dot{M}_{Edd} \\equiv L_{Edd}/c^2$, cooling becomes important as in spherical accretion (Park 1990). In such high $\\dot M/\\dot{M}_{Edd}$ case, the existence and the properties of the flow depend on the specific value of $\\rho_{out}/\\rho_0$. Figure 2 shows the mass accretion rate of constructed solutions as a function of the angular momentum at the outer boundary, in $\\dot{m}$ versus $\\lambda$ plane. Symbols represent the solutions for different $T_{out}$: circles for $T_{out}=1.1\\times10^{10} \\K$ ($r_{out}=2.5\\times 10^2 r_{Sch}$), triangles for $T_{out}=5.5\\times10^9 \\K$ ($r_{out}=5.0\\times 10^2 r_{Sch}$), squares for $T_{out}=2.8\\times10^9 \\K$ ($r_{out}=1.0\\times 10^3 r_{Sch}$), pentagons for $T_{out}=1.1\\times10^9 \\K$ ($r_{out}=2.5\\times 10^3 r_{Sch}$), and hexagons for $T_{out}=5.5\\times10^8 \\K$ ($r_{out}=5.0\\times 10^3 r_{Sch}$). All these solutions are calculated for $\\rho_{out} = 1.5 \\times 10^{-6} \\rho_0$. The solutions for $T_{out}=1.1\\times10^7 \\K$ ($r_{out}=2.5\\times 10^5 r_{Sch}$), temperature suitable for the ISM in a galactic nucleus, have much larger $r_{out}$, and the bifurcation between subsonic to unphysical branch becomes even sharper. Since the critical mass accretion rate above which the hot, optically thin accretion disk does not exist also decreases as $r_{out}$ increases (Abramowicz et al. 1995) and the Bondi rate increases as $T_{out}$ decreases, we choose $\\rho_{out} = 1.5 \\times 10^{-9} \\rho_0$ so that $\\dot{M} /\\dot{M}_{Edd}$ is in a range where solutions can be found. The mass accretion rates of these solutions are shown as crosses in the center of Figure 2. We find that regardless of $T_{out}$, the mass accretion rate $\\dot{m}$ decreases as the angular momentum $\\lambda$ increases. It approaches the Bondi accretion rate as $\\lambda$ decreases. This change of the mass accretion rate on the angular momentum is expected because given density and temperature at $r_{out}$, $\\rho$ and $H$ are fixed in the continuity equation (Eq. [\\ref{eq:continuity}]), and the mass accretion rate is determined solely by the radial velocity $v_r$. From equation (\\ref{eq:radial}), the radial velocity is determined by the difference between the gravity and the centrifugal accleration which is proportional to $l^2/r^3$ (at $r \\gg r_{Sch}$). Larger angular momentum causes smaller infall velocity, and the mass accretion rate decreases, and vice versa. The mass accretion rate for the lowest angular momentum flow is quite close to the Bondi rate ($\\dot{m} \\sim 1$) while that for the highest angular momentum flow is roughly 20 times smaller than the corresponding Bondi rate ($\\dot{m} \\sim 0.05$). We also find that $\\dot{m}$ as a function of $\\lambda$ is rather insensitive to $T_{out}$, or equally $r_{out}$ for $\\dot{m} \\ga 0.1$: The slope of $\\dot m(\\lambda)$ has a tendency to become slightly flatter (from circles to crosses) as $T_{out}$ decreases but the difference is not large. Since we expect the flow profile, especially the radial velocity, to depend on the specific value of $\\alpha$, we also calculate the solutions for different values of $\\alpha$. The mass accretion rates of $\\alpha=0.003$ are shown as a series of stars in the left side of Figure 2 and that of $\\alpha=0.03$ are shown in the right side of of Figure 2. Compared to the solutions for $\\alpha=0.01$ in the center denoted by crosses, lower $\\alpha$ flows have smaller mass accretion rate and higher $\\alpha$ flows larger mass accretion rate. Larger $\\alpha$ means stronger viscosity for given density and temperature, which causes larger radial velocity, and hence larger mass accretion rate, and vice versa for smaller $\\alpha$. The relation between $\\dot m$ and $\\lambda$ for $0.1 \\la \\dot{m} \\la 1$ can be approximated by \\begin{equation}\\label{eq:mdot_mB} \\dot{m}(\\lambda) = 0.09 \\left( \\frac{\\alpha}{0.01} \\right) \\lambda^{-1}. \\end{equation} The three straight lines (from left to right) in Figure 2 show this fitting function for $\\alpha=0.003$, 0.01, and 0.03, respectively. The mass accretion rate of the self-similar ADAF that has density $\\rho_{out}$ at $r_{out}$ has the same dependency on $\\rho_{out}$ and $r_{out}$ as the Bondi rate $\\dot{M}_{B}$ (Narayan \\& Yi 1994). Hence, the dimensionless mass accretion rate of self-similar ADAF is simply $\\dot{m}_{ADAF} = 3.3 \\alpha$. Since ADAF generally refers to $\\lambda \\sim 1$ flow and $\\dot{m}(\\lambda=1) \\simeq 3 \\dot{m}_{ADAF}$, we find that our solutions have approximately three times higher mass accretion rates than those of self-similar ADAFs. The difference is probably due to the fact that the density and velocity of global ADAFs do not exactly follow the self-similar form, especially near the outer boundary (Narayan, Kato, \\& Honma 1997; Chen, Abramowicz, \\& Lasota 1997). For given $T_{out}$, the lower limit on the mass accretion rate is determined by the Keplerian angular momentum barrier. A smaller mass accretion rate demands smaller radial infall velocity. From equation (\\ref{eq:angular}), smaller $|v_r|$ requires larger $\\Omega$, and eventually $\\Omega$ becomes larger than the Keplerian value $\\Omega_K$ at $r_{out}$ for too small mass accretion rate. Gas cannot accrete in steady-state when the angular momentum is much larger than the Keplerian one, and therefore no solution exists when $\\lambda \\ga 3$. Although we could determine $\\dot m$ below $\\dot{m} \\simeq 0.1$ from shooting until $\\lambda$ approaches $\\sim 1.5 - 3$ or 3, in many cases we failed to construct the full transonic solutions down to the inner boundary. The difficulty is much more severe for small $\\alpha$: the bifurcation between subsonic and unphysical occurs at a large radius and computing accuracy is not enough to integrate all the way down to the horizon. So we have limited confidence in the solutions below $\\dot{m} \\la 0.2$ for $\\alpha = 0.003$ and 0.03. The upper limit on the mass accretion rate on the other hand is determined by the Bondi accretion rate itself. The lowest angular momentum flow has its mass accretion rate already quite close to the Bondi rate, which is the mass accretion rate for zero angular momentum flow. Thus any higher mass accretion flow will end up on the unphysical branch of solutions in spherical Bondi flow (type III), with minimum effect due to the angular momentum. We can think of extending the velocity versus radius diagram of Bondi (1952) to a similar one with the angular momentum effect added. What angular momentum does is to lower the critical mass accretion rate for the transonic flow. As the angular momentum of the flow increases, the critical mass accretion rate decreases, and the rotating accretion flow accretes transonically at a critical mass accretion rate lower than the original Bondi accretion rate. On the other hand, if the angular momentum of the flow decreases, the critical mass accretion rate increases up to the Bondi accretion rate which is the critical mass accretion rate for zero angular momentum flow, above which no transonic steady-state rotating accretion flow exists. Since equation (\\ref{eq:mdot_mB}) gives $\\dot{m} > 1$ for $\\lambda < 9 \\alpha$ and $\\dot{m}$ cannot be greater than unity, it looks like there is no solution for $\\lambda < 9 \\alpha$. However, we are using slim disk formulation with a simple viscosity prescription to describe what is really a two-dimensional quasi-spherical flow, especially for lowest angular momentum flow; so the non-existence of solutions below the lower limit on $\\lambda$ is probably caused by the approximate physical descriptions of the flow, especially the overestimated viscosity for the low or no angular momentum flow. Surely, we will have a spherical Bondi flow when the angular momentum is zero. Therefore, we expect that steady flow probably exists even below $\\lambda \\sim 9 \\alpha$ and \\begin{equation}\\label{eq:mdot_zerolambda} \\dot{m} \\simeq 1 \\quad {\\rm for} \\quad \\lambda \\la 9 \\alpha . \\end{equation} We conclude that the steady-state, transonic hot accretion flow exists only when the mass accretion rate is within a certain range or the angular momentum at the outer boundary is below a certain maximum value. For our specific choice of parameters and conditions, the range in the mass accretion rate is \\begin{equation}\\label{eq:mdot_range} \\dot{m}_{cr} \\la \\dot{m} \\leq 1, \\end{equation} where $\\dot{m}_{cr}=\\dot{m}(\\lambda_{cr})$ is the mass accretion rate of the flow with maximum possible angular momentum at the outer boundary. Our calculations suggest $\\lambda_{cr} \\la 3$ and $\\dot{m}_{cr} \\sim 0.05$ for $\\alpha = 0.01$. Now we discuss flows with the outer boundary temperature different from the virial temperature. For example, when the cool, outer geometrically thin disk is attached to the hot inner accretion disk, the temperature at the outer boundary will be much lower than the virial temperature (Narayan, Kato, \\& Honma 1997). So we fix $r_{out} = 10^3 ~r_{Sch}$ but vary $T_{out}$. The outer boundary density is the same at $\\rho_{out} = 1.5 \\times 10^{-6} \\rho_0$. Figure 3 shows the mass accretion rate again in $\\dot{m}$ versus $\\lambda$ plane. Symbols are the same as in Figure 2, except circles for $T_{out}=3.6\\times10^9 \\K$, triangles for $T_{out}=2.2\\times10^9 \\K$, squares for $T_{out}=1.1\\times10^9 \\K$, pentagons for $T_{out}=5.5\\times10^8 \\K$, and hexagons for $T_{out}=2.0\\times10^8 \\K$. Again, solutions exist only in a limited range of $\\dot{m}$, or equally $\\lambda$. But this time both upper and lower limits on $\\dot{m}$ vary with $T_{out}$. The reason for the difference is that $r_{out}$ is fixed and not related to $T_{out}$ whereas it is related to $T_{out}$ in Bondi solutions. As in previous choice of boundary, the lower limit on $\\dot{m}$ is from the Keplerian angular momentum barrier. However, the upper limit is determined for a different reason in this case. As the mass accretion rate for given boundary conditions increases, the radial infall velocity at the outer boundary should increase. But the increasing radial infall velocity eventually becomes larger than the sound speed at the boundary. The whole flow becomes supersonic from the outer boundary to the innermost radius. Since we are looking for transonic solutions that uniquely determine the mass accretion rate, we disregard these supersonic branch of solutions, and no transonic solutions exist above certain upper limit on $\\dot{m}$. This upper limit on $\\dot{m}$ decreases as $T_{out}$ decreases because the sound speed decreases with $T_{out}$ and the flow becomes supersonic at smaller radial velocity, or mass accretion rate, causing rapid decrease of the upper limit on $\\dot{m}$ with decreasing $T_{out}$. The dependence of $\\dot{m}$ on $\\lambda$ is also different. Again, $\\dot{m}$ decreases as $\\lambda$ increases, but is not inversely proportional to $\\lambda$, nor is independent of $T_{out}$. Cooler flow, i.e., lower $T_{out}$, has lower mass accretion rate $\\dot{m}$ for a given $\\lambda$. When $\\lambda = 0.1$, $\\dot{m}$ for $T_{out}=2.2\\times10^9 \\K$ is $\\sim 1$ while that for $T_{out}=2.0\\times10^8 \\K$ is as low as $\\sim 10^{-4}$. This means that if gas is injected into a given radius with near-Keplerian angular momentum but with temperature much below the virial temperature at that radius, the mass accretion rate can be a few orders of magnitude below the Bondi accretion rate." }, "0910/0910.0117_arXiv.txt": { "abstract": "We use Monte-Carlo Markov chain techniques to constrain acceptable parameter regions for the Munich L-Galaxies semi-analytic galaxy formation model. Feedback from active galactic nuclei (AGN) is required to limit star-formation in the most massive galaxies. However, we show that the introduction of tidal stripping of dwarf galaxies as they fall into and merge with their host systems can lead to a reduction in the required degree of AGN feedback. In addition, the new model correctly reproduces both the metallicity of large galaxies and the fraction of intracluster light. ", "introduction": "This paper describes the implementation of a model for dwarf galaxy disruption within the semi-analytic framework of \\citet[hereafter DLB07]{LuB07}. The original model was suggested by \\citet{HBT08} as a way of both reducing the excess of dwarf galaxies and creating the intracluster light (ICL); however this was implemented \\emph{a posteriori}, acting only to reduce the dwarf population at the current day. The new model follows the stripping of dwarfs as they fall into the halos of their parent galaxy, thus gradually reducing their mass: affecting their infall rates, increasing the time-scale for, and decreasing the magnitude of, the merger with the central object. For the purposes of these conference proceedings, the most important result is that the masses of the black holes are reduced, with a corresponding reduction in the level of feedback of AGN energy into the interstellar medium. This should be read in conjunction with the paper by Chris Short in this volume that investigates the degree to which the accretion energy is needed to provide feedback into the intracluster medium (ICM) in order to provide the observed entropy excess in clusters. ", "conclusions": "The black hole growth model presented here provides a real challenge for models that require AGN heating to raise the entropy of the ICM. Even using the original DLB07 model, with its higher black hole masses, 35 per cent of the available rest mass energy is required (see article by Chris Short in this volume, also \\citet{ShT09}). Similar conclusions have been reached by \\citet{BMB08} using the Durham GALFORM semi-analytic model. \\begin{theacknowledgments} This work was undertaken using the Virgo Consortium cluster of computers, COSMA. BH acknowledges the support of his PhD scholarship from the Portuguese Science and Technology Foundation which supported him for most of the time while this work was developed. PAT was supported by an STFC rolling grant. \\end{theacknowledgments}" }, "0910/0910.5218_arXiv.txt": { "abstract": "We calculate the cosmic microwave background (CMB) bispectrum due to inhomogeneous reionization. We calculate all the terms that can contribute to the bispectrum that are products of first order terms on all scales in conformal Newtonian gauge. We also correctly account for the de-correlation between the matter density and initial conditions using perturbation theory up to third order. We find that the bispectrum is of local type as expected. For a reasonable model of reionization, in which the universe is completely ionized by redshift $z_{ri}\\sim 8$ with optical depth to the last scattering surface $\\tau_0=0.087$ the signal-to-noise ratio for detection of the CMB temperature bispectrum is $\\rm{S/N}\\sim 0.1$ and confusion in the estimation of primordial non-Gaussianity is $f_{NL}\\sim -0.1$. For an extreme model with $z_{ri}\\sim 12.5$ and $\\tau_0=0.14$ we get $\\rm{S/N}\\sim 0.5$ and $f_{NL}\\sim -0.2$. ", "introduction": " ", "conclusions": "" }, "0910/0910.2865_arXiv.txt": { "abstract": "Densely-packed, all-digital aperture arrays form a key area of technology development required for the Square Kilometre Array (SKA) radio telescope. The design of real-time signal processing systems for digital aperture arrays is currently a central challenge in pathfinder projects worldwide. We describe an hierarchical, frequency-domain beamforming architecture for synthesising a sky beam from the wideband antenna feeds of digital aperture arrays. ", "introduction": "Densely-packed, all-digital aperture arrays form a key area of technology development required for the Square Kilometre Array (SKA) radio telescope. The design of real-time signal processing systems for aperture arrays is certainly one of the most challenging tasks in the design of next-generation instruments. In this paper, we describe the design and implementation of an hierarchical beamforming architecture for wideband radio telescopes.\\\\ This document is organised as follows. Chapter I is an introduction to the problem domain. Chapter II describes the theory and process of `beamforming' for aperture arrays used in radio astronomy. Chapter III describes the design and implementation of a frequency-domain beamformer. Chapter IV describes calibration procedures for the array, and is followed by results and conclusions.\\\\ \\subsection{Aperture Phased Arrays} An aperture phased array is a collection of fixed, non-directional antennas whose outputs are combined to synthesise an effective directional antenna. This synthesised antenna is pointed electronically; significantly the physical component antennas do not ever move. The electronic combination of signals with beamforming techniques performs the same spatial filtering task as a parabolic dish would in a traditional radio telescope dish. Figure \\ref{BeamformingIllustration} illustrates the basic principle of this operation.\\\\ The flexibility, survey speed, system-level autonomy and scalability provided by antenna arrays make them extremely attractive for next-generation radio instruments, for use, in our application, in radio astronomy, but also applicable to other wideband detection and communication instruments. Future radio instruments will certainly be dependent on the availability of high-performance signal processing systems such as these. Since aperture arrays are seen as a critical antenna technology to achieve the required sky survey speed in future radio telescopes (see, for example, \\cite{2006SKA.memo.81} and \\cite{2007skads.memo.FeA}), the development of a scalable signal processing system is of great importance.\\\\ \\begin{figure} \\begin{center} \\includegraphics[width=70mm]{BeamformingDiagNoBackg.pdf} \\caption {An illustration of the electronic processing of signals received from an array of antennas for concentrating the signal from a certain direction. This is the basic principle of the phased aperture array: different delay configurations allow beams to be formed in different directions of \\(\\theta\\)} \\label{BeamformingIllustration} \\end{center} \\end{figure} ", "conclusions": "In this paper we propose that digital aperture arrays are a fast, useful and elegant collector technology for GHz radio astronomy, that the development of high-performance signal processing systems is critical to the success of next-generation radio instruments, and have described the design of a 4-element frequency-domain aperture array beamformer.\\\\" }, "0910/0910.4391_arXiv.txt": { "abstract": "If dark energy (DE) couples to neutrinos, then there may be apparent violations of Lorentz/{\\it CPT} invariance in neutrino oscillations. The DE-induced Lorentz/{\\it CPT} violation takes a specific form that introduces neutrino oscillations that are energy independent, differ for particles and antiparticles, and can lead to novel effects for neutrinos propagating through matter. We show that ultra-high-energy neutrinos may provide one avenue to seek this type of Lorentz/{\\it CPT} violation in $\\nu_\\mu$-$\\nu_\\tau$ oscillations, improving the current sensitivity to such effects by seven orders of magnitude. Lorentz/{\\it CPT} violation in electron-neutrino oscillations may be probed with the zenith-angle dependence for high-energy atmospheric neutrinos. The ``smoking gun,'' for DE-neutrino coupling would, however, be a dependence of neutrino oscillations on the direction of the neutrino momentum relative to our peculiar velocity with respect to the CMB rest frame. While the amplitude of this directional dependence is expected to be small, it may nevertheless be worth seeking in current data and may be a target for future neutrino experiments. ", "introduction": "The accelerated cosmic expansion \\cite{SNIa} poses difficult questions for theoretical physics~\\cite{Copeland:2006wr,Caldwell:2009ix,Silvestri:2009hh}. Is it simply due to a cosmological constant? Is some new negative-pressure dark energy (DE) required? Is general relativity modified at large distance scales? The major thrust of the empirical assault on these questions has been to determine whether the expansion history and growth of large-scale structure are consistent with a cosmological constant or require something more exotic \\cite{DETF}. However, it may be profitable to explore whether there are other experimental consequences of the new physics---which we collectively refer to as DE, although it may involve a modification of gravity rather than the introduction of some new substance---responsible for accelerated expansion. If cosmic acceleration is due to a cosmological constant (i.e., if general relativity is valid and the equation-of-state parameter is $w=-1$), then the vacuum is Lorentz invariant. If, however, something else is going on, then the ``vacuum'' has a preferred frame: the rest frame of the cosmic microwave background (CMB). If, moreover, dark energy couples somehow to standard-model particles, then there may be testable (apparent) violations of Lorentz invariance. For example, if DE is coupled to the pseudoscalar $F\\tilde F$ of electromagnetism \\cite{Carroll:1998zi}, there may be a ``cosmological birefringence'' that rotates the linear polarization of cosmological photons; CMB searches for such a rotation \\cite{Lue:1998mq} constrain this rotation to be less than a few degrees~\\cite{Feng:2006dp}. Here we explore DE-induced Lorentz/{\\it CPT}-violating effects in the neutrino sector. We show that the form of a Lorentz-violating coupling between neutrinos and dark energy is highly restricted under fairly general assumptions.\\footnote{The coupling of neutrinos to dark energy has also been considered in the context of ``mass-varying neutrinos''~\\cite{MaVaN}, but that implementation of the DE-neutrino coupling does not lead to the type of Lorentz/{\\it CPT}-violating effects we discuss here.} The coupling engenders an additional source for neutrino mixing (e.g., Ref.~\\cite{Gu:2005eq}), resulting in neutrino oscillations with a different energy dependence than vacuum oscillations and different oscillation probabilities for neutrinos and antineutrinos. While similar Lorentz/{\\it CPT}-violating oscillations have been considered before~\\cite{bargeretal,CPT_nu,Kostelecky2004}, we emphasize here that cosmic acceleration dictates a specific form for such effects. Data from Super-Kamiokande and K2K~\\cite{Gonzalez-Garcia2004} and AMANDA/IceCube~\\cite{Abbasi2009} already tightly constrain {\\it CPT}-violating parameters for $\\nu_\\mu$-$\\nu_\\tau$ mixing, and those from solar-neutrino experiments and KamLAND~\\cite{Bahcall2002} do so for $\\nu_e$-$\\nu_\\mu$ mixing. However, the effects of DE-induced {\\it CPT} violation become more significant at higher energies \\cite{Hooper:2005}. Here we show that next-generation measurements of ultra-high-energy neutrinos produced by spallation of ultra-high-energy cosmic rays will increase the sensitivity to {\\it CPT}-violating $\\nu_\\mu$-$\\nu_\\tau$ oscillations by seven orders of magnitude. We also show that these {\\it CPT}-violating couplings may lead to novel effects in the zenith-angle dependence for atmospheric neutrinos in the $\\sim100$~GeV range. While such {\\it CPT}-violating effects, if detected, could be attributed simply to intrinsic {\\it CPT} violation in fundamental physics, not related to DE, a DE-neutrino coupling further predicts a directional effect: the neutrino-mixing parameters depend on the neutrino propagation direction relative to our peculiar velocity with respect to the CMB rest frame. While this signature will likely remain elusive even to next-generation experiments, it would, if detected, be a ``smoking gun'' for DE beyond a cosmological constant. It is therefore worth considering as a long-range target for future neutrino experiments. It may also be worthwhile to search current data in case an implementation of DE-neutrino coupling different from that we consider here leads to a different energy dependence for these directional effects. We therefore work out explicitly the directional dependence to aid experimentalists who may wish to look for such correlations in current data. Below, we first derive in Sec.~\\ref{sec:formalism} the form of the Lorentz/{\\it CPT} violation allowed by a DE-neutrino coupling and discuss the resulting neutrino-oscillation physics. In Sec.~\\ref{sec:cosmogenic} we apply the formalism to cosmogenic ultra-high-energy neutrinos, and obtain projected sensitivities of future detectors to these effects in $\\nu_\\mu$-$\\nu_\\tau$ oscillations. In Sec.~\\ref{resonance} we discuss matter-induced effects for $\\nu_e$ oscillations in high-energy atmospheric neutrinos in the presence of Lorentz-invariance--violating mixings. Concrete formulas for the directional dependence on oscillation probabilities are given in Sec.~\\ref{sec:direction}. Finally, we discuss some theoretical implications in Sec.~\\ref{implications} and summarize and conclude in Sec.~\\ref{sec:conclusion}. ", "conclusions": "\\label{sec:conclusion} We studied the implications of an interaction between dark energy and neutrinos for neutrino oscillations. The most general Lorentz/{\\it CPT}-violating term induced by dark energy (DE) takes the form $(a_L)^\\mu \\bar\\nu \\gamma_\\mu (1-\\gamma_5) \\nu$, where $(a_L)^\\mu$ is a four-vector normal to the CMB rest frame. This introduces a new source for neutrino oscillations that are energy independent and different for neutrinos and antineutrinos. Furthermore, the motion of the Earth with respect to the cosmic rest frame induces a directional dependence in the oscillation probabilities. The current best limits to the DE-neutrino coupling we considered are obtained from atmospheric- and accelerator-neutrino experiments for $\\nu_\\mu$-$\\nu_\\tau$ mixing, and from solar and reactor experiments for $\\nu_e$-$\\nu_\\mu$ mixing. However, the higher the neutrino energy, the more prominent the effect of the DE-neutrino interaction. We therefore considered in this paper cosmogenic ultra-high-energy (energies of $10^{17}$--$10^{19}$ eV) neutrinos produced by the interaction of ultra-high-energy cosmic rays with CMB photons. We showed that future experiments targeting these neutrinos will improve the sensitivity to a DE-neutrino interaction by seven orders of magnitude, down to $m_{\\rm eff} \\sim 10^{-30}$ GeV compared with the current upper bound $m_{\\rm eff} \\alt 5\\times 10^{-23}$ GeV (Fig.~\\ref{fig:GZK}). This corresponds to a sensitivity to an energy scale as large as $\\sim$10$^6$ GeV for the DE-neutrino interaction. We then showed that the interplay of DE- and matter-induced neutrino mixing could induce a novel zenith-angle dependence for $\\nu_e$ oscillations in atmospheric neutrinos. This effect may extend the sensitivity to Lorentz/{\\it CPT}-violating parameters in the $\\nu_e$ by roughly three orders of magnitude. The real smoking gun of a DE-neutrino interaction (as opposed to some other origin for Lorentz/{\\it CPT} violation) would be a directional dependence of the oscillation probabilities. The notion that Lorentz violation may give rise to a directional dependence is not new (e.g., Ref.~\\cite{Kostelecky:2004hg}) and searches for directional dependence in neutrino experiments have already been carried out (e.g., Ref.~\\cite{Adamson:2008ij}), but prior work has considered Lorentz-violating parameters introduced in an {\\it ad hoc} manner and/or tested for direction-dependent effects in a Sun-centered inertial frame. We emphasize here that cosmic acceleration suggests that we seek a specific form of Lorentz violation, that where the preferred frame is aligned with the CMB rest frame. Even though such a signal is expected to be small, it is still worth seeking in existing and future experimental data. We have not discussed specific models for a DE-neutrino interaction, beyond an illustrative toy model, but it may be interesting to do so (see also Ref.~\\cite{stephon}). The theoretical motivation to expect such a coupling may admittedly be slim. However, we are at square one in our understanding of DE, and such a coupling is no less likely to be expected, perhaps, than any of the many other manifestations of new cosmic-acceleration physics that have been considered. Discovery of Lorentz/{\\it CPT}-violating effects would be extremely important, even if not attributable directly to dark energy. A directional dependence, if discovered, would be absolutely remarkable, as it would provide moreover clear evidence that there is more to cosmic acceleration than simply a cosmological constant." }, "0910/0910.1327_arXiv.txt": { "abstract": "Spectral evolution models are a widely used tool for determining the stellar content of galaxies. I provide a review of the latest developments in stellar atmosphere and evolution models, with an emphasis on massive stars. In contrast to the situation for low- and intermediate-mass stars, the current main challenge for spectral synthesis models are the uncertainties and rapid revision of current stellar evolution models. Spectral libraries, in particular those drawn from theoretical model atmospheres for hot stars, are relatively mature and can complement empirical templates for larger parameter space coverage. I introduce a new ultraviolet spectral library based on theoretical radiation-hydrodynamic atmospheres for hot massive stars. Application of this library to star-forming galaxies at high redshift, i.e., Lyman-break galaxies, will provide new insights into the abundances, initial mass function and ages of stars in the very early universe. ", "introduction": "Spectral evolution models attempt to reproduce the observed spectra of systems ranging in sizes from star clusters to luminous galaxies by numerically combining models or observations of star formation, stellar evolution, and stellar spectra. \\cite[Tinsley (1968)]{Tinsley68} is generally credited for initiating the field, and \\cite[Charlot \\& Bruzual's (1991)]{ChaBru91} introduction of the isochrone synthesis technique validated this approach for ages up to a Hubble time. The success of synthetic spectral models echoes the increased availability of computers in astronomy. The power of this technique has been amplified over the past decade by the growing importance of the internet and databases for storage, query, and distribution of spectral templates and related models. This trend is revolutionizing the field in a way similar to Tinsley's and Charlot's \\& Bruzual's achievements. In this talk I will concentrate less on the technical but more on some of the astrophysical aspects of spectral evolution models. After a brief discussion of the main ingredients in the model I will address the two major components: stellar evolution models and stellar libraries. The emphasis will be on {\\em massive stars}, for which stellar evolution still is the major challenge, whereas reliable stellar libraries are becoming available in larger and larger numbers. This situation is somewhat opposite to the case of low-mass stars. I will introduce a new ultraviolet (UV) spectral library constructed from a grid of radiation-hydrodynamic models and discuss its potential for interpreting the rest-frame UV spectra of Lyman-break galaxies. Detecting the first stellar generations and modeling and understanding their spectra will be a major effort for the next decade. ", "conclusions": "" }, "0910/0910.4502_arXiv.txt": { "abstract": "We investigate the viscous two temperature accretion disc flows around rotating black holes. We describe the global solution of accretion flows with a sub-Keplerian angular momentum profile, by solving the underlying conservation equations including explicit cooling processes selfconsistently. Bremsstrahlung, synchrotron and inverse Comptonization of soft photons are considered as possible cooling mechanisms. We focus on the set of solutions for sub-Eddington, Eddington and super-Eddington mass accretion rates around Schwarzschild and Kerr black holes with a Kerr parameter $0.998$. It is found that the flow, during its infall from the Keplerian to sub-Keplerian transition region to the black hole event horizon, passes through various phases of advection -- general advective paradigm to radiatively inefficient phase and vice versa. Hence the flow governs much lower electron temperature $\\sim 10^8-10^{9.5}$K, in the range of accretion rate in Eddington units $0.01\\lsim\\mdot\\lsim 100$, compared to the hot protons of temperature $\\sim 10^{10.2}-10^{11.8}$K. Therefore, the solution may potentially explain the hard X-rays and $\\gamma$-rays emitted from AGNs and X-ray binaries. We then compare the solutions for two different regimes of viscosity and conclude that a weakly viscous flow is expected to be cooling dominated, particularly at the inner region of the disc, compared to its highly viscous counter part which is radiatively inefficient. With all the solutions in hand, we finally reproduce the observed luminosities of the under-fed AGNs and quasars (e.g. Sgr~$A^*$) to ultra-luminous X-ray sources (e.g. SS433), at different combinations of input parameters such as mass accretion rate, ratio of specific heats. The set of solutions also predicts appropriately the luminosity observed in the highly luminous AGNs and ultra-luminous quasars (e.g. PKS~0743-67). ", "introduction": "The cool Keplerian accretion disc (Pringle \\& Rees 1972; Shakura \\& Sunyaev 1973; Novikov \\& Thorne 1973) was found to be inappropriate to explain observed hard X-rays, e.g. from Cyg~X-1 (Lightman \\& Shapiro 1975). It was argued that secular instability of the cool disc swells the optically thick, radiation dominated region to a hot, optically thin, gas dominated region resulting in hard component of spectrum $\\sim 100$KeV (Thorne \\& Price 1975; Shapiro, Lightman \\& Eardley 1976). This region is strictly of two temperatures with electron and ion temperatures respectively $\\sim 10^9$K and $\\sim 5\\times 10^{11}$K which confirms that cool, one temperature, pure Keplerian accretion solution is not unique. Indeed Eardley \\& Lightman (1975) found that a Keplerian disc is unstable due to thermal and viscous effects when viscosity parameter $\\alpha$ (Shakura \\& Sunyaev 1973) is constant. Later Eggum et al. (1985) showed by numerical simulations that the Keplerian disc with a constant $\\alpha$ collapses. Around eighties, therefore, the idea of two component accretion disc started floating around. For example, Paczy\\'nski \\& Wiita (1980) described a geometrically thick regime of the accretion disc in the optically thick limit, while Rees et al. (1982) introduced accretion torus in the optically thin limit. Moreover the idea of sub-Keplerian, transonic accretion was introduced by Muchotrzeb \\& Paczy\\'nski (1982), which was later improved by other authors (Chakrabarti 1989, 1996; Mukhopadhyay 2003). Other models were proposed by e.g. Gierli\\'nski et al. (1999), Coppi (1999), Zdziarski et al. (2001), including a secondary component in the accretion disc. On the other hand, Narayan \\& Yi (1995) introduced a two temperature disc model in the regime of inefficient cooling resulting in a vertical thickening of the hot disc gas. Here the pressure forces are expected to become important in modifying the disc dynamics which is likely to be sub-Keplerian. Other models with similar properties were proposed by, e.g., Begelman (1978), Liang \\& Thompson (1980), Rees et al. (1982), Eggum, Coroniti \\& Katz (1988). Abramowicz et al. (1988) proposed a height-integrated disc model, namely ``slim disc\", having high optical depth of the accreting gas at super-Eddington accretion rate such that the diffusion time is longer than the viscous time. The model was further applied to study the thermal and viscous instabilities in optically thick accretion discs (Wallinder 1991; Chen \\& Taam 1993). Shapiro, Lightman \\& Eardley (1976) initiated a two temperature Keplerian accretion disc at a low mass accretion rate which is optically thin and significantly hotter than the single temperature Keplerian disc of Shakura \\& Sunyaev (1973). The optically thin hot gas cools down through the bremsstrahlung and inverse-Compton processes and could explain various states of Cyg~X-1 (Melia \\& Misra 1993). Similarly, the ``ion torus\" model by Rees et al. (1982) was applied to explain AGNs at a low mass accretion rate. However, the two temperature model solutions by Shapiro, Lightman \\& Eardley (1976) appear thermally unstable. Narayan \\& Popham (1993) and subsequently Narayan \\& Yi (1995) showed that introduction of advection may stabilize the system. However, the solutions of Narayan \\& Yi (1995), while of two temperatures, could explain only a particular class of hot systems with inefficient cooling mechanisms. They also described the hot flow based on the assumption of ``self similarity\" which is just a ``plausible choice\". They kept the electron heating decoupled from the disc hydrodynamical computations which merely is an assumption. Later on, the solutions were attempted to generalize by Nakamura et al. (1997), Manmoto et al. (1997), Medvedev \\& Narayan (2001), relaxing efficiency of cooling into the systems, but concentrating only on specific classes of solutions. On the other hand, Chakrabarti \\& Titarchuk (1995) and later Mandal \\& Chakrabarti (2005) modeled two temperature accretion flows around Schwarzschild black holes in the general ``advective paradigm\", emphasizing possible formation of shock and its consequences therein. However, they also did not include the effect of electron heating self-consistently into the hydrodynamical equation, and thus the hydrodynamical quantities do not get coupled to the rate of electron heating (see also Rajesh \\& Mukhopadhyay 2009). In the present paper, we model a selfconsistent accretion flow in the regime of two temperature transonic sub-Keplerian disc (see also Sinha, Rajesh \\& Mukhopadhyay 2009; a brief version of the present work, but around Schwarzschild black holes). We consider all the hydrodynamical equations of the disc along with thermal components and solve the coupled set of equations selfconsistently. We neither restrict to the advection dominated regime nor the self-similar solutions. We allow the disc to cool selfconsistently according to the thermo-hydrodynamical evolution and compute the corresponding cooling efficiency factor as a function of radial coordinate. We investigate that when does the disc switch from the radiatively inefficient nature to general advective paradigm and vice versa. In order to implement our model to explain observed sources, we focus on the under-luminous AGNs and quasars (e.g. Sgr~$A^{*}$), ultra-luminous quasars and highly luminous AGNs (e.g. PKS~0743-67) and ultra-luminous X-ray (ULX) sources (e.g. SS433), when the last items are likely to be the ``radiation trapped'' accretion discs. While the first two cases correspond to respectively sub-Eddington and super-Eddington accretion flows around supermassive black holes, the last case corresponds to super-Eddington accretors around stellar mass black holes. In the next section, we discuss the model equations describing the system and the procedure to solve them. Subsequently, we discuss the two temperature accretion disc flows around stellar mass and supermassive black holes, respectively in \\S3 and \\S4, for both sub-Eddington, Eddington and super-Eddington accretion rates. Section 5 compares the disc flow of low Shakura-Sunyaev (1973) $\\alpha$ with that of high $\\alpha$ and then between the flows around co and counter rotating black holes. Then we discuss the implications of the results with a summary in \\S6. ", "conclusions": "We model the two temperature accretion flow, particularly around black holes, combining the equations of conservation and comprehensive cooling processes. We consider self-consistently the important cooling mechanisms: bremsstrahlung, synchrotron and inverse Comptonization due to synchrotron photons, where ions and electrons are allowed to have different temperatures. As matter falls in, hot electrons cool through the various cooling mechanisms, particularly by the synchrotron emission when the magnetic field is high. This is particularly the case for the flow around stellar mass black holes where the magnetic field may also act as a boost in transporting the angular momentum. However, in the present paper, we do not consider such processes in detail, rather stick with the standard $\\alpha$-prescription. By solving a complete set of disc equations we show that in general the disc system exhibits GAAF. However, in certain circumstances GAAF becomes radiatively inefficient, depending on the flow parameters and hence efficiency of cooling mechanisms. Transitions from GAAF to radiatively inefficient flow and vice versa are clearly explained by the cooling efficiency factor $f$, shown in each model cases. While the previous authors, who proposed ADAF (Narayan \\& Yi 1994, 1995), especially restricted with flows having $f=1$ (inefficient cooling), here we do not impose any restriction to the flow parameters to start with and let the parameter $f$ to determine self-consistently as the system evolves. Therefore, our model is very general whose special case may be understood as a radiatively inefficient advection dominated flow. We have explored especially the optically thin flows incorporating bremsstrahlung, synchrotron and inverse Comptonization processes. In Fig. \\ref{opti} we show the variation of the effective optical depth as a function of disc radii for two limiting cases. While flows around rotating black holes appear slightly thinner compared to the corresponding cases of static black holes, in general $\\tau_{\\rm eff}\\lsim 5\\times 10^{-4}$. This verifies our choice of optically thin flows throughout. However, for the present purpose, when the main aim is to understand disk dynamics in the global, viscous, two-temperature regime, we have ignored inverse Comptonization due to bremssstrahlung photons, if any. This may be important in cases of very super-Eddington accretion flows which we plan to explore in future, particularly, in analysing the underlying spectra. \\begin{figure} \\centering \\includegraphics[width=0.50\\columnwidth,angle=-90]{f22.eps} \\caption{ Variation of the effective optical depth as a function of radial coordinate. Solid ($a=0$) and dotted ($a=0.998$) curves correspond to $M=10$, $\\mdot=100$ and dashed ($a=0$) and dot-dashed ($a=0.998$) curves correspond to $M=10^7$, $\\mdot=0.01$. See Tables 1 and 2 for details. } \\label{opti} \\end{figure} The temperature of the flow depends on the accretion rate. If the accretion rate is low and thus the flow is radiatively inefficient, then the disc is hot. Such a hot flow is being attempted to model since 1976 (Shapiro, Lightman \\& Eardley 1976) when it was assumed that locally $Q^+\\sim Q^-$ and thus $f\\rightarrow 0$. While the model was successful in explaining observed hard X-rays from Cyg~X-1, it turned out to be thermally unstable. Rees et al. (1982) proposed a hot ion torus model avoiding $f$ to unity. In the similar spirit Narayan \\& Yi (1995) proposed the hot two temperature solution in the assumption of $f\\rightarrow 1$ including the strong advection into the flow. Abramowicz et al. (1995), based on the single temperature model, showed that the optically thin disc flow of accretion rate more than one Eddington does not have an equilibrium solution. However, they did not attempt to solve the complete set of differential equations. Based on some simplistic assumptions they showed the importance of advective cooling. Moreover, a single temperature description does not allow them to include all the underlying physics necessary to describe the cooling processes. In the due course, Mandal \\& Chakrabarti (2005) proposed a two temperature disc solution where the ion temperature could be as high as $\\sim 10^{12}$K. However, they particularly emphasized on how does the shock in the disc flow enable cooling through the synchrotron mechanism, without carrying out a complete analysis of the dynamics. The present paper describes, to our knowledge, the first comprehensive work to model the two temperature accretion flow self-consistently by solving the complete set of underlying equations without any pre-assumptive choice of the flow variables to start with. The generality lies not only in its construction but also its ability to explain the under-luminous to ultra-luminous sources, stellar mass to supermassive black holes. Table 3 lists the luminosities for a wide range of parameter sets, obtained by our model. It reveals that for a very low mass accretion rate $\\mdot=0.0001$ around a supermassive black hole, the luminosity comes out to be $L\\sim 10^{34}$ erg/sec, which indeed is similar to the observed luminosity from a under-luminous source Sgr~$A^{*}$. In other extreme, for $\\mdot=100$ around a similar black hole, $L\\sim 10^{47}$ erg/sec, similar to what observed from the highly luminous AGNs like PKS~0743-67. On the other hand, when the black hole is considered to be of stellar mass, then at a high $\\mdot=100$, the model reveals $L\\sim 10^{40}$ erg/sec which is similar to the observed luminosity from ULX sources (e.g. SS433). \\clearpage \\centerline{ Table 3: Luminosity in erg/sec} \\begin{center} \\begin{tabular}{lllllllllllll} \\hline \\hline $\\mdot$ & $M$ & $\\gamma$ & $L$ \\\\ \\hline \\hline $0.0001$ & $10^7$ & $1.6$ & $10^{34}$ \\\\ $0.01$ & $10^7$ & $1.5$ & $10^{38}$ \\\\ $1$ & $10^7$ & $1.35$ & $5\\times 10^{42}$\\\\ $100$ & $10^7$ & $1.34$ & $10^{47}$\\\\ \\hline \\hline $0.01$ & $10$ & $1.5$ & $10^{33}$\\\\ $1$ & $10$ & $1.35$ & $7\\times 10^{36}$\\\\ $100$ & $10$ & $1.34$ & $10^{40}$\\\\ \\hline \\hline \\end{tabular} \\end{center} In general, an increase of accretion rate increases density of the flow which may lead to a high rate of cooling and thus decrease of the cooling factor $f$. Hence, $f$ is higher, close to unity reassembling radiatively inefficient flows, for sub-Eddington accretors, and is lower, sometimes close to zero, for super-Eddington flows. Actual value of $f$ in a flow also depends on the behaviour of hydrodynamic variables which determine the rate of cooling processes. Naturally, as the flow advances from the transition region to the event horizon, $f$ varies between $0$ and $1$. However, if the black hole is considered to be rotating, the flow angular momentum decreases and thus the radial velocity increases. This in turn reduces the residence time of the sub-Keplerian flow hindering cooling processes to complete. This flow is then expected to be hotter and hence $f$ to be higher compared to that around a static black hole. Therefore, the system may tend to be radiatively inefficient, evenif its counter part around a static black hole appears to be an GAAF. However, this also depends on the value of $\\alpha$, as shown in Fig. \\ref{alfacom}. A low value of $\\alpha$ increases the residence time of matter in the disc which helps in cooling processes to complete, rendering a radiatively inefficient flow to switch over to GAAF. This feature may help in understanding the transient X-ray sources. In all the cases, the ion and electron temperatures merge or tend to merge at around transition radius. This is because, the electrons are in thermal equilibrium with the ions and thus virial around the transition radius, particularly when $\\mdot\\gsim 1$. As the sub-Keplerian flow advances, the ions become hotter and the corresponding temperature increases, rendering the ion-electron Coulomb collisions weaker. The electrons, on the other hand, cool down via processes like bremsstrahlung, synchrotron emissions etc. keeping the electron temperature roughly constant upto very inner disc. This reveals the two temperature flow strictly. Important point to note is that we have assumed throughout the coupling between the ions and electrons is due to the Coulomb scattering. However, the inclusion of possible nonthermal processes of transferring energy from the ions and electrons (Phinney 1981, Begelman \\& Chiueh 1988) might modify the results. However, as argued by Narayan \\& Yi (1995), the collective mechanism discussed by Begelman \\& Chiueh (1988) may dominate over the Coulomb coupling at either a very low $\\alpha$ or a very low $\\mdot$. Instead, the viscous heating rate of ions is much larger than the collective rate of nonthermal heating of electrons, unless $\\alpha$ is too small what we have not considered in the present cases. Therefore, the assumption to neglect nonthermal heating of electrons is justified. Now the future jobs should be to understand the radiation emitted by the flows discussed here and to model the corresponding spectra. This will be the ultimate test of the model in order to explain observed data. \\appendix" }, "0910/0910.1920_arXiv.txt": { "abstract": "We explore potential of current and next-generation \\gr\\ telescopes for the detection of weak magnetic fields in the intergalactic medium. We demonstrate that using two complementary techniques, observation of extended emission around point sources and observation of time delays in \\gr\\ flares, one would be able to probe most of the cosmologically and astrophysically interesting part of the \"magnetic field strength\" vs. \"correlation length\" parameter space. This implies that \\gr\\ observations with {\\it Fermi} and ground-based Cherenkov telescopes will allow to (a) strongly constrain theories of the origin of magnetic fields in galaxies and galaxy clusters and (b) discover, constrain or rule out the existence of weak primordial magnetic field generated at different stages of evolution of the Early Universe. ", "introduction": "Magnetic fields are known to play an important role in the physics of a variety of astrophysical objects, from stars to galaxies and galaxy clusters. The existence of galactic magnetic fields with strengths in the range $1-10\\ \\mu$G is established via observations of Faraday rotation and Zeeman splitting of atomic lines in the radio band and of polarization of starlight in the optical band \\cite{kulsrud,beck08}. Magnetic fields of similar strength are found in the cores of galaxy clusters \\cite{carilli02}. Weaker magnetic fields with strengths in the range $10^{-8}-10^{-7}$~G were recently discovered at the outskirts of galaxy clusters \\cite{xu06,kronberg07}. Although the strength and spatial structure of magnetic fields in the Milky Way and some other galaxies are reasonably well known today, there is no commonly accepted theory about the origin of these magnetic fields (see \\cite{kronberg94,grasso00,widrow02,kulsrud} for recent reviews of the subject). There is a general agreement that the observed microGauss magnetic fields are the result of amplification of weak \"seed\" fields. The amplification mechanisms under discussion are the so-called \"$\\alpha-\\omega$\" dynamo (in the case of spiral galaxies) and/or compression and turbulent motions of plasma during the galaxy/cluster formation processes. The nature of the initial weak seed fields for the dynamo or turbulent amplification is largely unknown \\cite{kronberg94,grasso00,widrow02,kulsrud}. It might be that the seed fields are produced during the epoch of galaxy formation by electrical currents generated by the plasma experiencing gravitational collapse within a proto-galaxy \\cite{pudritz89,gnedin00}, or ejected by the first supernovae \\cite{rees87} or active galactic nuclei \\cite{donnert09}. Otherwise, the seed fields might originate from still earlier epochs of the Universe expansion, down to the cosmological phase transitions or inflation times \\cite{grasso00}. Wide uncertainties in both the mechanism of amplification of the seed fields and in the nature of the seed fields themselves have led, over the last half-a-century, to the appearance of a long-standing problem of the \"origin of cosmic magnetic fields\" (in galaxies and galaxy clusters). It is clear that the clue for the solution of this problem might be given by the measurements of the initial seed fields. However, up to recently there was little hope that the extremely weak fields outside galaxies and galaxy clusters would ever be detected. In what follows we show that direct measurements of the seed fields and derivation of constraints on their nature become possible with the newly available observations in the very-high-energy \\gr\\ band with space and ground-based \\gr\\ telescopes such as {\\it Fermi}, HESS, MAGIC, VERITAS and, in the near future, CTA, AGIS and HAWC. The method of measurement of \"ExtraGalactic\" Magnetic Fields (EGMF) with \\gr\\ telescopes is based on the possibility of detection of emission from electromagnetic cascade initiated by the primary \\gr s emitted by an extragalactic source and developing throughout the InterGalactic Medium (IGM) along the line of sight toward the source \\cite{plaga,neronov07,japanese,elyiv09,kachelriess09}. Based on the knowledge of sensitivity of existing and future \\gr\\ telescopes, we find the range of EGMF parameters, such as the field strength $B$ and the correlation length $\\lambda_B$, in which the EGMF is accessible for the measurements with one of the two available measurement techniques (imaging \\cite{neronov07,elyiv09,kachelriess09} or timing \\cite{plaga,japanese} of the cascade signal). We demonstrate that most of the astrophysically and cosmologically interesting range of EGMF parameters could be probed with \\gr\\ observations. The plan of the paper is as follows. In Section \\ref{sec:experiment} we summarize the existing bounds on the strength and correlation length of EGMF which come mostly from radio observations. In Section \\ref{sec:cosmology} we discuss limits on the cosmological magnetic fields from cosmology. Then, in Section \\ref{sec:theory} we compare the existing bounds to the theoretical predictions of two classes of models (\"astrophysical\" vs. \"cosmological\" models) of the \"seed\" fields and show that model predictions normally fall largely below the existing bounds. In Sections \\ref{sec:gammaray} -- \\ref{sec:strongB} we first summarize the methods of measurement of EGMF with \\gr\\ telescopes and then estimate the ranges of EGMF parameters which can be probed with different observational techniques and different telescopes. Finally, in Section \\ref{sec:conclusions} we draw conclusions from our study. ", "conclusions": "\\label{sec:conclusions} We have discussed the prospects of detection of weak magnetic fields in the intergalactic medium with the novel techniques of timing and imaging observations with ground and space-based \\gr\\ telescopes. These techniques enable measurements of extremely weak magnetic fields with strengths much lower than the ones accessible for the measurements with radio telescopes (via Zeeman splitting and/or Faraday rotation techniques). We have demonstrated that using \\gr\\ observations one can detect, or rule out the possibility of existence of cosmologically or astrophysically produced \"seed\" magnetic fields in the voids of the large scale structure, which are conjectured to exist in a range of theories of the origin of magnetic fields in galaxies and galaxy clusters (see Fig. \\ref{fig:exclusion_theory}). To summarize, we find that discovery or non-detection of weak magnetic fields in the voids of the large scale structure should provide, in the nearest future, a decisive test of the theories of the origin of cosmic magnetic fields." }, "0910/0910.1934_arXiv.txt": { "abstract": "The dynamical properties of a model of dark energy in which two scalar fields are coupled by a noncanonical kinetic term are studied. We show that overall the addition of the coupling has only minor effects on the dynamics of the two-field system for both potentials studied, even preserving many of the features of the assisted quintessence scenario. The coupling of the kinetic terms enlarges the regions of stability of the critical points. When the potential is of an additive form, we find the kinetic coupling has an interesting effect on the dynamics of the fields as they approach the inflationary attractor, with the result that the combined equation of state of the scalar fields can approach $-1$ during the transition from a matter dominated universe to the recent period of acceleration. ", "introduction": "\\label{sec:Introduction} One goal of cosmology is to understand the origin of the observed accelerated expansion of the universe (see \\cite{Copeland:2006wr} and references therein). To date, there are several suggestions, including the cosmological constant, slowly evolving scalar fields and modifications to Einstein's theory of general relativity. Since scalar fields are predicted by many particle physics theories, scalar field models of dark energy, such as quintessence \\cite{Wetterich:1987fm, Ratra:1987rm, Martin:2008qp} or k-essence \\cite{Chiba:1999ka, ArmendarizPicon:2000dh, ArmendarizPicon:2000ah} have been studied in considerable depth in the past. One requirement a satisfactory model of dark energy must fulfill is that it leads to an equation of state (EOS) close to $w=-1$, in order to agree with current observational data. For single scalar fields, exponential potentials with slope $\\lambda$ lead to a scalar field dominated universe with late-time accelerated expansion if $\\lambda<\\sqrt{2}$. In this case, the duration of the matter dominated epoch depends on the initial conditions for the scalar field. Thus, the situation is not better than that with a cosmological constant. If, on the other hand, $\\lambda>\\sqrt{3(1+w)}$, the scalar field scales with the dominant fluid (with EOS $w$ \\cite{Ferreira:1997hj, Copeland:1997et}). This would help the initial condition problem, but unfortunately, the scaling solution and the accelerating solution are mutually exclusive. The situation with inverse power-law potentials is better in the sense that a wide range of initial conditions end up with the same cosmology at late times, but in order for the theory to be consistent with observational data, the exponent has to be small. On the other hand, if quintessence is indeed the scenario realized in nature, the quintessence sector might turn out to be rather nontrivial. There might, for example, be several scalar fields interacting \\cite{Barreiro:1999zs,Coley:1999mj,Blais:2004vt,Kim:2005ne} and/or the scalar fields might interact with matter in the universe \\cite{Amendola:1999er,TocchiniValentini:2001ty, Brookfield:2007au, Bean:2008ac}. In this paper we will study interacting scalar fields as a model for dark energy and study whether this can shed some light on the issues in quintessence model building. Models with multiple scalar fields have been considered in the past: partly to study isocurvature (entropy) perturbations (e.g. \\cite{Langlois:1999dw,Gordon:2000hv,Byrnes:2006fr,Langlois:2008vk}) and/or non-Gaussianity generated during inflation (see e.g. \\cite{Alishahiha:2004eh,Rigopoulos:2005ae,Hattori:2005ac,Cai:2009hw} and references therein). It was also found that the cumulative effect of many scalar fields could relax the constraints on the inflationary potential \\cite{Liddle:1998jc,Malik:1998gy,Calcagni:2007sb}. These ideas have also been used in models of dark energy (assisted quintessence \\cite{Kim:2005ne,Tsujikawa:2006mw,Ohashi:2009xw}). Like the case with inflation in the very early universe, there is no reason why dark energy is not driven by several interacting scalar fields. In a scenario with several scalar fields one might hope to find an explanation for the coincidence problem, i.e. why dark energy dominates today and not earlier during the cosmic history. In single scalar field models this is rather difficult to achieve (see e.g. \\cite{Copeland:2006wr}). We consider a simple extension of the standard case with canonical normalized fields by allowing for a cross-term in the kinetic energy of the scalar field, which, in the case of a homogeneous and isotropic universe, is proportional to $\\dot\\phi\\dot\\chi$. This is the simplest extension possible in the two-field case without changing the properties of the potential energy. Kinetic interactions between multiple scalar fields have been previously studied in the context of models in which the dark energy equation of state can become less than $-1$ \\cite{Sur:2009jg,Chimento:2008ws}. The paper is organized as follows: in the next section we present the model and discuss some aspects of the dynamics of the system, such as effective exponents. In Secs. \\ref{sec:assist} and \\ref{sec:soft} we perform a critical point analysis for two types of exponential potentials and in Sec. \\ref{sec:assist_num} we present the results of a numerical analysis. In Sec. \\ref{sec:var_a} we consider the case of a varying coupling function. Our conclusions can be found in Sec. \\ref{sec:discussion}. ", "conclusions": "\\label{sec:discussion} Although the use of scalar fields to model the acceleration of the universe could well be described as standard practice in cosmology, the fundamental physics that underpins these phenomenological models is by no means clear. It is wise, therefore, to consider the possibility that the observed acceleration of the universe may have more exotic origins. The recent focus on scalar fields with noncanonical kinetic terms derived from string-inspired models has shown that models that deviate from the standard lore are capable of producing fruitful results that display qualitatively different behaviour. In this work we have considered a model of dark energy in which the kinetic terms of two scalar fields are coupled by a term in the Lagrangian of the form $a(\\p_\\alpha\\phi\\p^\\alpha\\chi)$. This allowed us to study the simplest extension of the standard case without modifying the potential energy. After analysing the phase plane structure of the model for two different potentials, we found that the basic features of the uncoupled case were preserved and the most important effect of the extra degree of freedom was to change the range of parameters that lead to the different types of solutions. In particular, there exist stable points corresponding to scalar field-dominant or scaling solutions for every value of $\\lambda$, $\\mu$ and $a$ considered. In the case of the assisted potential we found that for negative $a$, critical points C1, D1, C2 and D2 that existed but were unstable in the case without the kinetic coupling became stable. This means that when $a$ goes from positive to negative values, there are three different regions in the $\\lambda,\\mu$-parameter space that give rise to accelerated solutions, each with a different value of the EOS, $w_{\\rm fields}$. This is less interesting from a phenomenological point of view, however, as except when the values of the potential exponents are very small, negative $a$ values have $w_{\\rm fields}$ larger than the observed dark energy EOS. In the case of positive $a$, one finds the EOS to be closer to $-1$ than that in the model with $a=0$, though in the region of interest ($w_{\\rm fields}\\approx-1$) the effect of the kinetic coupling is extremely slight. The soft potential evinced a similar dependence on $a$, though it is of less interest as a cosmological model as accelerated behaviour only occurs with small values of the exponents in all cases. Although models with exponential potentials exhibit both scaling and scalar field-dominant solutions, it is difficult to solve the coincidence problem without fine-tuning the initial conditions or allowing the parameters to vary in time. In Sec. \\ref{sec:assist_num} we discuss how the kinetic coupling affects the dynamics of the fields as they evolve towards the critical points. When $a$ is close to $2$ one finds that the value of $w_{\\rm fields}$ becomes very close to $-1$ for a brief period before the scalar fields dominate. This is a partial resolution of the perennial problem of the EOS being too large. This analysis is revealing as the addition of a cross-kinetic term has only a relatively minor effect on the dynamics of the two-field system for both cases studied, even preserving many of the features of the assisted quintessence scenario. In the case of the assisted potential, the kinetic coupling has an interesting effect on the dynamics of the fields as they approach the stable solution, with the result that the EOS of the scalar fields can approach $-1$ during the transition from a matter dominated universe to the recent period of acceleration, something that does not occur in assisted quintessence or the single field case. A natural extension to this work is to consider the case when the scalar fields are coupled directly to the matter fluid by a similar mechanism to that described in \\cite{TocchiniValentini:2001ty}. In this case, one would expect the value and stability ranges of the critical points to change dramatically due to the extra degree(s) of freedom, but it would be interesting to see whether the additive potential still allows large values of the potential exponents. The work could also be extended by treating $\\lambda$ and $\\mu$ as dynamical variables in a manner similar to \\cite{Ng:2001hs}, in which case it may be possible to obtain solutions with a large range of initial conditions that track the background solution and later dominate. As we have discussed, our results are also applicable in the case of a slowly varying coupling function $a(\\phi,\\chi)$ as well as the case in which the coupling undergoes a rapid change but is constant (or evolving very slowly) after the transition." }, "0910/0910.3877_arXiv.txt": { "abstract": "We present the near-infrared luminosity and stellar mass functions of galaxies in the core of the Shapley supercluster at $z$=0.048, based on new $K$-band observations carried out at the United Kingdom Infra-Red Telescope with the Wide Field Infrared Camera in conjunction with $B$- and $R$-band photometry from the Shapley Optical Survey, and including a subsample ($\\sim$650 galaxies) of spectroscopically confirmed supercluster members. These data sets allow us to investigate the supercluster galaxy population down to M$_K^\\star$+6 and $\\mathcal{M}$=10$^{8.75}$M$_\\odot$. For the overall 3\\,deg$^2$ field the $K$-band luminosity function (LF) is described by a Schechter function with M$_K^\\star$=--24.96$\\pm$0.10 and $\\alpha$=--1.42$\\pm$0.03, a significantly steeper faint-end slope than that observed in field regions. We investigate the effect of environment by deriving the LF in three regions selected according to the local galaxy density, and observe a significant ($2\\sigma$) increase in the faint-end slope going from the high- ($\\alpha$=--1.33) to the low-density ($\\alpha$=--1.49) environments, while a faint-end upturn at M$_{K}>$--21 becomes increasingly apparent in the lower density regions. The galaxy stellar mass function (SMF) is fitted well by a Schechter function with log$_{10}$($\\mathcal{M}^\\star$)=11.16$\\pm$0.04 and $\\alpha$=--1.20$\\pm0.02$. The SMF of supercluster galaxies is also characterised by an excess of massive galaxies that are associated to the brightest cluster galaxies. While the value of $\\mathcal{M}^*$ depends on environment increasing by 0.2 dex from low- to high-density regions, the slope of the galaxy SMF does not vary with the environment. By comparing our findings with cosmological simulations, we conclude that the environmental dependences of the LF are not primary due to variations in the merging histories, but to processes which are not treated in the semi-analytical models, such as tidal stripping or harassment. In field regions the SMF shows a sharp upturn below $\\mathcal{M}$=10$^{9}$M$_\\odot$, close to our mass limit, suggesting that the upturns seen in our $K$-band LFs, but not in the SMF, are due to this dwarf population. The environmental variations seen in the faint-end of the $K$-band LF suggests that these dwarf galaxies, which are easier to strip than their more massive counterparts, are affected by tidal/gas stripping upon entering the supercluster environment. ", "introduction": "\\label{intro} The properties and evolution of galaxies are strongly related to their environment (e.g. Blanton et al. \\citeyear{bla05a}; Rines et al. \\citeyear{rin05}; Baldry et al. \\citeyear{BBB06}), through the mass and merging histories of their host dark matter halos, and the impact of different physical mechanisms (e.g. Treu et al. \\citeyear{tre03}) that are linked in various ways to the local galaxy density and the properties of the intergalactic medium. In the local Universe this environmental dependence has been investigated and observed in the distribution of galaxy luminosities and stellar masses, providing constraints on the assembly of galaxies over cosmic time (see below). Since the NIR light is dominated by established old stellar populations rather than by recent star-formation activity, the NIR LF can be considered as reliable estimator of the stellar mass function (SMF, Gavazzi et al. \\citeyear{GPB96}; Bell \\& de Jong \\citeyear{BdJ01}) and the shape of the NIR LF constrains the scenarios of galaxy formation (e.g. Benson et al. \\citeyear{BBF03}), with different energetic feedback processes from supernovae and AGN required to simultaneously fit the LFs at the faint and bright ends respectively \\citep{BBM06}. Early versions of the hierarchical galaxy formation model predicted a decrease of the abundance of massive galaxies with redshift (Kauffmann \\& Charlot \\citeyear{KC98}) which have continued forming until recent times through processes of merging and accretion, and a steep mass function due to the presence of a large number of faint dwarf galaxies witnessing the small dark halo formation in the early universe (e.g. Kauffmann et al. \\citeyear{KWG93}). The NIR LFs observed at different redshifts turn out to be well described by Schechter functions, although there is some evidence for an excess of bright galaxies with respect to the best fitting Schechter function (e.g. Jones et al. \\citeyear{JPC06}), while a faint-end upturn has also been observed in some cases (e.g. De Propris \\& Pritchet \\citeyear{DPP98}; Balogh et al. \\citeyear{BCZ01}; Jenkins et al. \\citeyear{JHM07}). The results supporting the hierarchical scenario (e.g. Kauffmann \\& Charlot \\citeyear{KC98}) are contrasted by other works that bring it into question (Kodama \\& Bower \\citeyear{KB03}; De Propris et al. \\citeyear{DSE99}; \\citeyear{DSE07}) or cannot interpret univocally their findings (e.g. Drory et al. \\citeyear{DBF03}). For instance, there is a consensus that the evolution of the characteristic absolute luminosity M$^\\star$ for both field and cluster galaxies can be described by a passively evolving population formed in a single burst at redshift $z$=1.5-2 (e.g. Lin et al. \\citeyear{LMG06}; Wake et al. \\citeyear{WNE06}; De Propris et al. \\citeyear{DSE99}; \\citeyear{DSE07}). On the other hand, Drory et al. (\\citeyear{DBF03}) attribute the observed evolution in the $K$-band LF of cluster galaxies, either to a change in the mass-to-light ratio alone (i.e. passive evolution), or to a combination of changes in M/L and stellar mass which could be due to star formation and/or to merging or accretion. Recent semi-analytic models incorporating AGN feedback have been able to reproduce well the evolution of the SMF over $0{<}z{<}5$ (Bower et al. \\citeyear{BBM06}), predicting the observed population of massive galaxies at $z{>}2$. However, in order to properly use the NIR, LF to disentangle between possible scenarios of galaxy evolution, one has to consider the dependence of the LF on environment, galaxy colour, spectral type and morphology. The LF of field galaxies at NIR wavebands has been primarily investigated through the Two Micron All Sky Survey\\footnote{http://www.ipac.caltech.edu/2mass/} (2MASS) often complemented by either optical photometry from the Sloan Digital Sky Survey\\footnote{http://www.sdss.org/} (SDSS) or wide-field spectroscopic surveys such as the 2dF Galaxy Redshift Survey\\footnote{http://www.mso.anu.edu.au/2dFGRS/} (2dFGRS). These works (e.g. Kochanek et al. \\citeyear{KPF01}; Cole et al. \\citeyear{CNB01}) agree in describing the NIR LFs with Schechter functions characterised by a rather flat faint-end slope $\\alpha$ from --0.77 to --0.96 and which are independent of the morphological type (Kochanek et al. \\citeyear{KPF01}), in contrast with that found in optical surveys where the faint-end slope is steeper for late-type galaxies. On the contrary, Bell et al. (\\citeyear{BMK03}) found that NIR LF has a brighter characteristic luminosity and shallower slope for early-type galaxies with respect to the later types. More recently, Jones et al. (\\citeyear{JPC06}), using the 6dF Galaxy Redshift Survey\\footnote{http://www.aao.gov.au/local/www/6df/} (6dFGRS, Jones et al. \\citeyear{JSC04}) derived NIR LFs for field galaxies 1-2\\,mag deeper in absolute magnitude with respect to the previous works. They found that the Schechter function is not ideal to reproduce the data since it cannot match the bright- and faint-end simultaneously, due to an excess of galaxies at magnitudes brighter than M$^\\star$. \\begin{figure*} \\centerline{\\resizebox{17.5cm}{!}{\\includegraphics{fig1.eps}}} \\caption{The Shapley NIR survey (thick red) is shown superimposed to the surface density of $\\mathrm{R}<21.0$ galaxies as derived from the SOS (see text). Isodensity contours are shown at intervals of 0.25 galaxies arcmin$^{-2}$, with the thick contours corresponding to 0.5, 1.0 and 1.5 galaxies arcmin$^{-2}$, the densities used to separate the three cluster environments. The multi-wavelength photometric coverage of ACCESS on the SSC are also shown. Magenta: 24\\,$\\mu$m, green: 70\\,$\\mu$m, blue dashed: 150\\,nm and 250\\,nm.} \\label{fig1} \\end{figure*} Balogh et al. (\\citeyear{BCZ01}) investigated the dependence of the infrared galaxy luminosity function and the associated galaxy SMF on environment and spectral type by means of 2MASS and Las Campanas Redshift Survey (LCRS, Shectman et al. \\citeyear{SLO96}) for galaxies brighter that M$_J$=--19\\,mag. In field environments the LF of galaxies with emission lines turns out to have a much steeper faint-end slope ($\\alpha$=--1.39) compared to that of galaxies without emission lines ($\\alpha$=--0.59). On the other hand, in the cluster environment, even the non-emission line galaxies have a steep faint-end LF ($\\alpha$=--1.22). This difference is almost entirely due to the non-emission line galaxies which dominate the cluster population, and present a slope close to that of the overall field. Thus, they suggested that the cluster population is built up by accreting field galaxies with little effect other than the cessation of star formation. Differences in the shape of the LF for late- and early-type cluster galaxies has been found by Huang et al. (\\citeyear{HGC03}): the late-type galaxies having a systematically fainter M* and steeper faint-end slope. A possible faint-end upturn in the $H$-band LF was already suggested by De Propris et al. (\\citeyear{DES98}, see also Andreon \\& Pell\\'o \\citeyear{AP00}) for the Coma cluster outlining the steep trend of dwarf galaxies down to M*+5, even if they did not provide a precise estimate of the faint-end slope because of possible field contaminations. An increase of the faint-end slope of Coma was recently observed at 3.6\\,$\\mu$m by Jenkins et al. (\\citeyear{JHM07}) indicating a large number of faint red galaxies. However, Rines \\& Geller (\\citeyear{rin08}) found no such upturn for Virgo, based on a fully spectroscopically confirmed sample, and suggested that many of the photometrically-selected red sequence galaxies which contribute to the upturns seen in other clusters are background galaxies. Finally, the tight correlation between the total galaxy NIR luminosity and the cluster binding mass (Lin et al. \\citeyear{LMS03}, \\citeyear{LMS04}; Ramella et al. \\citeyear{RBG04}) allows to probe that the global cluster $K$-band mass-to-light ratio decreases with cluster radius (Rines et al. \\citeyear{RGD04}) showing that the environment affects the shape of the LF also within the clusters. We note that most of the previous works are based on the 2MASS data which has a detection sensitivity (10$\\sigma$) of $K$=13.1\\,mag for extended sources (Cole et al. \\citeyear{CNB01}, $K$=13.57\\,mag according to Bell et al. \\citeyear{BMK03}), limiting studies of the environmental impact on the NIR LF to only a sample of local clusters (e.g. Virgo), or limiting to magnitudes $<$M$^\\star$+2 (e.g. Rines et al. \\citeyear{RGD04}). However, the dominant processes that quench star-formation, and therefore transform galaxies, depend crucially on the galaxy mass (e.g. Haines et al. \\citeyear{HLM06}; \\citeyear{HGL07}), and the strong bimodality in the properties of galaxies about a characteristic stellar mass of $\\sim 3\\times 10^{10} M_{\\odot}$ ($\\sim$ M*+1, Kauffmann et al. \\citeyear{KHW03}) implies fundamental differences in the formation and evolution of giant and dwarf galaxies (e.g. Dekel \\& Birnboim \\citeyear{DB06}; Kere\\^s et al. \\citeyear{KKW05}). This issue needs data-sets reaching much fainter luminosities than those of 2MASS to be investigated in order to obtain, in general, an overall picture of galaxy evolution and, in particular, to establish the contribution of the dwarf galaxy population to the total stellar mass in the local universe and the physical origin of the claimed faint-end upturn. The recent development of wide-field NIR imagers on 4-m class telescopes such as UKIRT/WFCAM and KPNO/NEWFIRM has opened the possibility of NIR surveys to be $>100\\times$ more sensitive covering many square degrees (e.g. UKIDSS), allowing the $K$-band LF of nearby clusters to be obtained covering not only the cluster cores, but the entire virialized regions. In this context we study the $K$-band LF of the Shapley supercluster core (SSC) down to the dwarf regime (reaching $\\sim$M$^\\star$+6) with the aim of i) quantifying the environmental impact on the shape of the NIR LF; ii) deriving the stellar masses of the supercluster galaxies; iii) investigating the mechanisms driving galaxy evolution as function of galaxy mass. This work is carried out in the framework of the joint research programme ACCESS aimed at determining the importance of cluster assembly processes in driving the evolution of galaxies as a function of galaxy mass and environment within the Shapley supercluster (see Sect.~\\ref{ACCESS}). In Sect.~\\ref{sec:3}, we describe the data-sets. In Sect.~\\ref{sec:4}, we derive the NIR galaxy luminosity functions obtained through background subtraction in the whole observed field and we study the ongoing effects of environment by comparing the LFs of galaxies in three different regions of the supercluster, characterized by high-, intermediate- and low-density, we also compare NIR and optical LFs. The galaxy stellar mass function is presented in Sect.~\\ref{sec:5}. The results are discussed in Sect.~\\ref{sec:6} and the summary and conclusions of this work is given in Sect.~\\ref{sec:7}. Throughout the paper we adopt a cosmology with $\\Omega_M$=0.3, $\\Omega_\\Lambda$= 0.7, and H$_0$=70 km s$^{-1}$Mpc$^{-1}$. According to this cosmology 1 arcmin corresponds to 60 kpc at $z$=0.048 and the distance modulus is 36.66. ", "conclusions": "\\label{sec:7} It is well known that the NIR luminosities provide a more reliable estimate of the stellar masses compared to the optical ones due to the fact that the mass-to-light ratios in the NIR vary by at most a factor 2 across a wide range of star formation histories (Bell \\& de Jong \\citeyear{BdJ01}) in comparison to the much larger variations of the M/L ratios (up to a factor 10) observed at optical wavelengths. In addition the effects of the extinction are much weaker at infrared wavelengths that in the optical ones, and k-corrections for infrared colours are only weakly dependent on the Hubble type and vary slowly with redshift (e.g. Poggianti \\citeyear{pog97}). In our study we exploit new deep ($K$=18\\,mag) $K$-band imaging of the SSC complemented by the deep optical imaging down to $B$=22.5\\,mag and $R$=22\\,mag from the SOS, and a spectroscopically confirmed supercluster sample of $\\sim$650 galaxies across the same field which is $\\sim$90\\% complete at $R<$16\\,mag. We present an analysis of the $K$-band LF of galaxies as a function environment and we derive the galaxy SMF in order to constrain the physical mechanisms that transform galaxies in different environments as function of galaxy mass. Our results are summarized as follows. \\begin{description} \\item[-] The $K$-band LF can be fitted by a single Schechter function, with $M_{K}^{*}=-24.96\\pm0.10$ and $\\alpha=-1.42\\pm0.03$ in agreement with previous works of comparable depth. \\item[-] The $K$-band LF faint-end slope becomes steeper from high- to low-density environments varying from --1.33 to --1.49, being inconsistent at the 2$\\sigma$ c.l. (see Tab~\\ref{fitsLF}), indicating that the faint galaxy population increases in low-density environments. Such an environmental dependence confirms our finding for the optical LFs derived in the same supercluster regions although the changes in slope are less dramatic at NIR wavebands. \\item[-] The observed trend of the galaxy SMF presents a slope in agreement with previous works concerning the SMF of field galaxies (Cole et al. \\citeyear{CNB01}; Panter et al. \\citeyear{pan04}). The value of $\\mathcal{M}^\\star$ is higher with respect to that observed in the field (e.g. Bell et al. \\citeyear{BMK03}), as expected for cluster environment. The SMF of supercluster galaxies is characterised by an excess of massive galaxies that is associated to the cluster BCGs. Discrepancies with previous work that observed a strong faint-end upturn (Baldry et al. \\citeyear{BGD08} and Jenkins et al. \\citeyear{JHM07}) can be related to the different mass ranges investigated and/or environmental differences in the analysed structures. \\item[-] Differently from the LF no environment effect is found in the slope of the SMFs. On the other hand, the $\\mathcal{M}^\\star$ increase from low- to high-density regions and the excess of galaxies at the bright-end is also dependent on the environment. This trend is in general agreement with the results of Baldry et al. (\\citeyear{BBB06}). \\end{description} In order to interpret our findings, we use the Millennium simulation which produce DM halo and galaxy catalogues based on SAMs. The cluster NIR LFs obtained using the simulated catalogues do not show any significant variation with cluster-centric radius, thus suggesting that the variations observed in the LFs of the Shapley supercluster are not driven by variations in the merging histories of the galaxies, but are likely related to processes such as tidal stripping or harassment of infalling galaxies. By comparing the effect of environment at optical and NIR wavebands in shaping the LFs and taking into account that the slope of the galaxy SMF is invariant with respect to the environment, it seems that the physical mechanism responsible for the transformation of galaxies properties in different environment are related to the quenching of the star formation rather than mass-loss. This suggests that the mechanism responsible for shaping the LF and SMF is partially related to the mass loss due to tidal stripping or galaxy harassment, but gas stripping by the ICM can also affect the galaxy population removing the cold gas supply and thus rapidly terminating ongoing star-formation. These mechanisms all require a dense ICM and so their evolutionary effects on massive galaxies are limited to the cores of clusters, but can extend to poorer environments for dwarf galaxies which are easier to strip. The infalling galaxies are probably late-type (see MMH06) that are affected by gas stripping entering in the supercluster. On the other hand, the depth of the NIR survey could affect the present results by not allowing us to detect an upturn of the SMF which can be detected at a mass range lower than reached here (see Jenkins et al. \\citeyear{JHM07}). In order to identify the mechanisms which drive galaxy evolution in the supercluster environment we are undertaking a survey with the integral field spectrograph WiFeS which will provide a unique data set to investigate in details the stellar populations and kinematics for a subsample of the Shapley galaxies." }, "0910/0910.3793_arXiv.txt": { "abstract": "We examine deep \\emph{XMM-Newton} Reflection Grating Spectrometer (RGS) spectra from the cores of three X-ray bright cool core galaxy clusters, Abell~262, Abell~3581 and HCG~62. Each of the RGS spectra show Fe~\\textsc{xvii} emission lines indicating the presence of gas around 0.5~keV. There is no evidence for O~\\textsc{vii} emission which would imply gas at still cooler temperatures. The range in detected gas temperature in these objects is a factor of 3.7, 5.6 and 2 for Abell~262, Abell~3581 and HCG~62, respectively. The coolest detected gas only has a volume filling fraction of 6 and 3 per~cent for Abell 262 and Abell 3581, but is likely to be volume filling in HCG~62. \\emph{Chandra} spatially resolved spectroscopy confirms the low volume filling fractions of the cool gas in Abell~262 and Abell~3581, indicating this cool gas exists as cold blobs. Any volume heating mechanism aiming to prevent cooling would overheat the surroundings of the cool gas by a factor of 4. If the gas is radiatively cooling below 0.5~keV, it is cooling at a rate at least an order of magnitude below that at higher temperatures in Abell~262 and Abell~3581 and two-orders of magnitude lower in HCG~62. The gas may be cooling non-radiatively through mixing in these cool blobs, where the energy released by cooling is emitted in the infrared. We find very good agreement between smooth particle inference modelling of the cluster and conventional spectral fitting. Comparing the temperature distribution from this analysis with that expected in a cooling flow, there appears to be a even larger break below 0.5 keV as compared with previous empirical descriptions of the deviations of cooling flow models. ", "introduction": "The hot (few $10^7$K) intracluster medium (ICM) in galaxy clusters is primarily seen in emission in the X-ray waveband. The ICM contains most of the baryonic cluster mass. In a large fraction of clusters of galaxies, observations show that the X-ray surface brightness steeply rises towards their centres (e.g. \\citealt{Stewart84}). In addition, spectrally-derived temperature profiles of clusters typically show drops in temperatures by a factor of 2 or 3 from the outskirts (e.g. \\citealt{AllenSchmidtFabian01}). As the X-ray surface brightness is proportional to the integral of the density squared along a line of sight, peaked surface brightness profiles imply that the gas in the cores is cooling much more rapidly than in the outskirts. In the centre, the mean radiative cooling time, often calculated as the ratio of the luminosity of a region to its enthalpy, often drops below 1~Gyr. Assuming steady-state and the absence of heating, there will be material with short cooling times in the cluster centre cooling out of the X-ray emitting band. As the volume of this material becomes much smaller, there must be a flow of material in to replace it to preserve the steady state. The steep surface brightness profiles imply rates of 10s to 1000s of solar masses per year cooling out of the X-ray band (see \\citealt{Fabian94} for a review). This cooling material would be expected to eventually give rise to star formation. This picture changed when high spectral resolution X-ray studies of nearby clusters of galaxies using the Reflection Grating Spectrometer (RGS) instruments on \\emph{XMM-Newton} revealed a lack of cool X-ray emitting gas in these objects \\citep{Tamura01b, Tamura01a, Peterson01, Kaastra01, Sakelliou02, Peterson03}. The main spectral indicator missing in these spectra are the emission lines of Fe~\\textsc{xvii}, which are strong from material between 0.15 and 0.8~keV (1.7 to 9.4~MK). The general accepted picture is that there is a lack of material below a factor 2 or 3 of the outer temperature, which matches the results from \\emph{Chandra} spatially-resolved temperature profiles. This is not completely correct, however. Deep observations of nearby clusters show a much larger range of temperature. Fe~\\textsc{xvii} emission lines have been seen in Centaurus \\citep{SandersRGS08}, showing a temperature range of more than 10, Abell 2204 \\citep{SandersA220409}, showing a range of 15, 2A~0335+096 \\citep{Sanders2A033509}, showing a range of 8. Fe~\\textsc{xvii} emission was also detected in Abell~262, M87 and NGC~533 in a large study by \\cite{Peterson03}. There is much less material than would be expected to be seen in the case of steady-state radiative cooling without any heating, however. There is evidence that the central active galactic nuclei (AGN) in these objects can energetically prevent much of the cooling (see reviews by \\citealt{PetersonFabian06} and \\citealt{McNamaraNulsen07}). This heating may be by the inflation of cavities (radio lobes) by the AGN or the subsequent dissipation of sound waves generated by the inflation. Such cavities are found in almost all nearby clusters which require heating \\citep{DunnFabian06}. If AGNs are responsible for preventing cooling in cluster cores, there are still remaining issues. In Centaurus the cooling time of the lowest detected component is only $10^7$~yr \\citep{SandersRGS08}. If no cooling is taking place, then feedback must be able to operate on these timescales. Feedback must also be able to operate over the much longer cluster lifetime. Star formation in Centaurus must have occured slowly over the last 8~Gyr, or all must have been at earlier times \\citep{SandersEnrich06}. We assume a Hubble constant of $70 \\kmpspMpc$ and use the relative Solar metallicities of \\cite{AndersGrevesse89}. \\begin{table*} \\caption{Details of the targets and individual \\emph{XMM-Newton} observations. The exposure times given are for the RGS1 instrument after cleaning. The absorption quoted is the Galactic absorption from \\protect\\cite{Kalberla05}. References in superscript for redshifts are (1) \\protect\\cite{StrubleRood99}, (2) \\protect\\cite{Johnstone98} and (3) \\protect\\cite{ZabludoffMulchaey00}.} \\begin{tabular}{llllll} \\hline Cluster & Redshift & Absorption & Observation & Date & Exposure \\\\ & & ($10^{20}\\psqcm$) & & & (ks)\\\\ \\hline Abell 262 &$0.0163^1$& 5.67 & 0109980101 & 2001-01-16 & 26.2 \\\\ & & & 0504780101 & 2007-07-12 & 121.2 \\\\ & & & 0504780201 & 2007-07-18 & 41.4 \\\\ & & & total & & 188.8 \\\\ Abell 3581 &$0.0218^2$& 4.36 & 0205990101 & 2004-01-29 & 43.5 \\\\ & & & 0504780301 & 2007-08-01 & 116.3 \\\\ & & & 0504780401 & 2007-08-03 & 27.7 \\\\ & & & total & & 187.5 \\\\ HCG 62 &$0.01453^3$&3.31 & 0112270701 & 2003-01-15 & 12.4 \\\\ & & & 0504780501 & 2007-06-26 & 110.1 \\\\ & & & 0504780601 & 2007-06-29 & 33.6 \\\\ & & & total & & 156.6 \\\\ \\hline \\end{tabular} \\label{tab:sample} \\end{table*} \\begin{figure*} \\includegraphics[width=0.8\\textwidth]{fig01.pdf} \\caption{\\emph{Chandra} images of the cores of the three objects examined in this paper. Images have been smoothed with a 1 arcsec Gaussian. These images show the \\emph{Chandra} observations in Table~\\ref{tab:chandraobs}.} \\label{fig:chandraimages} \\end{figure*} \\begin{figure*} \\centering \\includegraphics[width=0.7\\textwidth]{fig02.pdf} \\caption{(Top left panel) Deprojected densities measured from \\emph{Chandra} spectral fitting (shown as points) and from surface brightness deprojection (shown as solid lines). (Top right panel) Deprojected temperature measured from \\emph{Chandra} spectral fitting (shown as points) and surface brightness deprojection (shown as lines). (Bottom left panel) Cooling time profiles measured from surface brightness deprojection (shown as small points) and RGS (shown as large points). (Bottom right panel) Cumulative mass deposition rates derived from \\emph{Chandra} surface brightness deprojection.} \\label{fig:chandraprofiles} \\end{figure*} \\begin{table} \\caption{\\emph{Chandra} datasets analysed.} \\begin{tabular}{llll} \\hline Target & Obs-ID & Date & Clean exposure (ks)\\\\ \\hline Abell 262 & 7921 & 2006-11-20 & 110.7 \\\\ Abell 3581 & 1650 & 2001-06-07 & 7.2 \\\\ HCG 62 & 921 & 2000-01-25 & 48.4 \\\\ \\hline \\end{tabular} \\label{tab:chandraobs} \\end{table} ", "conclusions": "" }, "0910/0910.5726_arXiv.txt": { "abstract": "We analyze the $>$100 MeV data for 3 GRBs detected by the {\\it Fermi} satellite (GRBs 080916C, 090510, 090902B) and find that these photons were generated via synchrotron emission in the external forward shock. We arrive at this conclusion by four different methods as follows. (1) We check the light curve and spectral behavior of the $>$100 MeV data, and late time X-ray and optical data, and find them consistent with the so called closure relations for the external forward shock radiation. (2) We calculate the expected external forward shock synchrotron flux at 100 MeV, which is essentially a function of the total energy in the burst alone, and it matches the observed flux value. (3) We determine the external forward shock model parameters using the $>$100 MeV data (a very large phase space of parameters is allowed by the high energy data alone), and for each point in the allowed parameter space we calculate the expected X-ray and optical fluxes at late times (hours to days after the burst) and find these to be in good agreement with the observed data for the entire parameter space allowed by the $>$100 MeV data. (4) We calculate the external forward shock model parameters using only the late time X-ray, optical and radio data and from these estimate the expected flux at $>$100 MeV at the end of the sub-MeV burst (and at subsequent times) and find that to be entirely consistent with the high energy data obtained by {\\it Fermi}/LAT. The ability of a simple external forward shock, with two empirical parameters (total burst energy and energy in electrons) and two free parameters (circum-stellar density and energy in magnetic fields), to fit the entire data from the end of the burst (1--50s) to about a week, covering more than eight-decades in photon frequency --- $>$10$^2$MeV, X-ray, optical and radio --- provides compelling confirmation of the external forward shock synchrotron origin of the $>$100 MeV radiation from these {\\it Fermi} GRBs. Moreover, the parameters determined in points (3) and (4) show that the magnetic field required in these GRBs is consistent with shock-compressed magnetic field in the circum-stellar medium with pre-shocked values of a few tens of micro-Gauss. ", "introduction": "The {\\it Fermi} Satellite has opened a new and sensitive window in the study of GRBs (gamma-ray bursts); for a general review of GRBs see Gehrels, Ramirez-Ruiz \\& Fox (2009), M\\'esz\\'aros (2006), Piran (2004), Woosley \\& Bloom (2006), Zhang (2007). So far, in about one year of operation, {\\it Fermi} has detected 12 GRBs with photons with energies $>$100 MeV. The $>$10$^2$ MeV emission of most bursts detected by the LAT (Large Area Telescope: energy coverage 20 MeV to $>$300 GeV) instrument aboard the {\\it Fermi} satellite shows two very interesting features (Omedei et al. 2009): (1) The first $>$100 MeV photon arrives later than the first lower energy photon ($\\lae$1 MeV) detected by GBM (Gamma-ray Burst Monitor), (2) The $>$100 MeV emission lasts for much longer time compared to the burst duration in the sub-MeV band (the light curve in sub-MeV band declines very rapidly). \\begin{table*} \\begin{center} \\begin{tabular}{ccccccc} \\hline & $\\beta_{LAT}$ & $p$ & $z$ & $d_{L28}$ & $t_{GRB}[s]$ & $E_{\\gamma,iso}[erg]$ \\\\ \\hline \\hline GRB 080916C & $1.20 \\pm 0.03$ & $2.4 \\pm 0.06$ & $4.3$ & $12.3$ & $60$ & $8.8\\times 10^{54}$\\\\ GRB 090510 & $1.1 \\pm 0.1$ & $2.2 \\pm 0.2$ & $0.9$ & $1.8$ & $0.3$ & $1.08\\times10^{53}$\\\\ GRB 090902B & $1.1 \\pm 0.1$ & $2.2 \\pm 0.2$ & $1.8$ & $4.3$ & $30$ & $3.63\\times10^{54}$\\\\ \\hline \\end{tabular} \\end{center} \\caption{{\\small The main quantities used in our analysis for these 3 GRBs. $\\beta_{LAT}$ is the spectral index for the $>100$MeV data, $p$ is the power-law index for the energy distribution of injected electrons i.e. $dn/d\\gamma\\propto \\gamma^{-p}$, $z$ is the redshift, $d_{L28}$ is the luminosity distance in units of $10^{28}$cm, $t_{GRB}$ is the approximate burst duration in the {\\it Fermi}/GBM band and $E_{\\gamma, iso}$ is the isotropic equivalent of energy observed in $\\gamma$-rays in the 10keV-10GeV band for GRB080916C and GRB090902B, and in the 10keV-30GeV band for GRB090510. Data taken from Abdo et al. (2009a, 2009b, 2009c), De Pasquale et al. (2010).}} \\end{table*} There are many possible $>$100 MeV photons generation mechanisms proposed in the context of GRBs; see Gupta \\& Zhang (2007) and Fan \\& Piran (2008) for a review. Shortly after the observations of GRB 080916C (Abdo et al. 2009a), we proposed a simple idea: the $>$100 MeV photons in GRB 080916C are produced via synchrotron emission in the external forward shock (Kumar \\& Barniol Duran 2009). This proposal naturally explains the observed delay in the peak of the light curve for $>$100 MeV photons -- it corresponds to the deceleration time-scale of the relativistic ejecta -- and also the long lasting $>$100 MeV emission, which corresponds to the power-law decay nature of the external forward shock (ES) emission (the ES model was first proposed by Rees \\& M\\'esz\\'aros 1992, M\\'esz\\'aros \\& Rees 1993, Paczy\\'nski \\& Rhoads 1993; for a comprehensive review of the ES model, see, e.g., Piran, 2004, and references therein). Following our initial analysis on GRB 080916C, a number of groups have provided evidence for the external forward shock origin of {\\it Fermi}/LAT observations (Gao et al. 2009; Ghirlanda, Ghisellini, Nava 2010; Ghisellini, Ghirlanda, Nava 2010; De Pasquale et al. 2010). In this paper we analyze the $>$100 MeV emission of GRB 090510 and GRB 090902B in detail, and discuss the main results of our prior calculation for GRB 080916C (Kumar \\& Barniol Duran 2009), to show that the high energy radiation for all these three arose in the external forward shock via the synchrotron process. These three bursts - one short and two long GRBs - are selected in this work because the high energy data for these bursts have been published by the {\\it Fermi} team as well as the fact that they have good afterglow follow up observations in the X-ray and optical bands (and also the radio band for GRB 090902B) to allow for a thorough analysis of data covering more than a factor 10$^8$ in frequency and $>10^4$ in time to piece together the high energy photon generation mechanism, and cross check this in multiple different ways. In the next section (\\S2) we provide a simple analysis of the LAT spectrum and light curve for these three bursts to show that the data are consistent with the external forward shock model. This analysis consists of verifying whether the temporal decay index and the spectral index satisfy the relation expected for the ES emission (closure relation), and comparing the observed flux in the LAT band with the prediction of the ES model (according to this model the high energy flux is a function of blast wave energy, independent of the unknown circum-stellar medium density, and extremely weakly dependent on the energy fraction in magnetic fields). We describe in \\S3 how the $>$100 MeV data alone can be used to theoretically estimate the emission at late times ($t\\gae$ a few hours) in the X-ray and optical bands within the framework of the external forward shock model, and that for these three bursts the expected flux according to the ES model is in agreement with the observed data in these bands. Moreover, if we determine the ES parameters ($\\epsilon_e$, $\\epsilon_B$, $n$, and $E$)\\footnote{$\\epsilon_e$ and $\\epsilon_B$ are the energy fraction in electrons and magnetic field for the shocked fluid, $n$ is the number density of protons in the burst circum-stellar medium, and $E$ is the kinetic energy in the ES blast wave.} using only the late time X-ray and optical fluxes (and radio data), we can predict the flux at $>$100 MeV at any time after the deceleration time for the GRB relativistic outflow. We show in \\S3 that this predicted flux at $>10^2$MeV is consistent with the value observed by the {\\it Fermi} satellite for the bursts analyzed in this paper. These exercises and results show that the high energy emission is due to the external shock as discussed in \\S3. We also describe in \\S3 that the magnetic field in the shocked fluid --- responsible for the generation of $>$ 100 MeV photons as well as the late time X-ray and optical photons via the synchrotron mechanism --- is consistent with the shock compression of a circum-stellar magnetic field of a few tens of micro-Gauss. It is important to point out that we do not consider in this work the prompt sub-MeV emission mechanism for GRBs --- which is well known to have a separate and distinct origin as evidenced by the very rapid decay of sub-MeV flux observed by Swift and {\\it Fermi}/GBM (the flux in the sub-MeV band drops-off with time as $\\sim t^{-3}$ or faster as opposed to the $\\sim t^{-1}$ observed in the LAT band). Nor do we investigate the emission process for photons in the LAT band during the prompt burst phase. ", "conclusions": "The {\\it Fermi} Satellite has detected 12 GRBs with $>$100 MeV emission in about one year of operation. In this paper we have analyzed the $>$100 MeV emission of three of them: two long-GRBs (090902B and 080916C) and one short burst (GRB 090510), and find that the data for all three bursts are consistent with synchrotron emission in the external forward shock. This idea was initially proposed in our previous work on GRB 080916C (Kumar \\& Barniol Duran 2009), shortly after the publication of this burst's data by Abdo et al. (2009a). Now, there are three GRBs for which high energy data has been published, and for all of them we have presented here multiple lines of evidence that $>$100 MeV photons, subsequent to the prompt GRB phase, were generated in the external forward shock. The reason that high energy photons are detected from only a small fraction of GRBs observed by {\\it Fermi} is likely due to the fact that the high energy flux from the external forward shock has a strong dependence on the GRB jet Lorentz factor, and therefore very bright bursts with large Lorentz factors are the only ones detected by {\\it Fermi}/LAT (this was pointed out by Kumar \\& Barniol Duran 2009, who also suggested that there should be no difference in long and short bursts, as far as the $>$100 MeV emission is concerned - the high energy flux is only a function of burst energy and time, eq. 1). We have analyzed the data in 4 different ways, and all of them lead to the same conclusion regarding the origin of $>$10$^2$MeV photons. First, we verified that the temporal decay index for the $>$100 MeV light curve and the spectral index are consistent with the closure relation expected for the synchrotron emission in the external forward shock. Second, we calculated the expected magnitude of the synchrotron flux at 100 MeV according to the external forward shock model and find that to be consistent with the observed value. Third, using the $>$100 MeV data only, we determined the external shock parameters, and from these parameters we predict the X-ray and optical fluxes at late times and find that these predicted fluxes are consistent with the observed values within the uncertainty of our calculations, i.e. a factor of two (see figs. \\ref{090902B-fwd}, \\ref{080916C-fwd}). And lastly, using the late time X-ray, optical and radio fluxes --- which the GRB community has believed for a long time to be produced in the external forward shock --- we determine the external shock parameters, and using these parameters we predict the expected $>$100 MeV flux at early times and find the flux to be in agreement with the observed value (see fig. \\ref{090902B-reverse}). The fact that the $>$100 MeV emission and the lower energy ($\\lae 1$ MeV) emission are produced by two different sources at two different locations suggests that we should be cautious when using the highest observed photon energy and pair-production arguments to determine the Lorentz factor of the GRB jet. We point out that the external shocks for these bursts were nearly adiabatic, i.e. radiative losses are small. The evidence for this comes from two different observations: (1) the late time X-ray spectrum lies in the adiabatic regime; (2) a radiative shock at early times (close to the deceleration time) would produce emission in the 10--10$^2$ keV band far in excess of the observed limits. We find that radiative shock is not needed to explain the temporal decay index of the $>$100 MeV light curve as suggested by (Ghisellini, Ghirlanda, Nava 2010), provided that the observing band is above all synchrotron characteristic frequencies. We find that the magnetic field required in the external forward shock for the observed high and low energy emissions for these three bursts is consistent with shock-compressed magnetic field in the CSM; the magnetic field in the CSM -- before shock compression -- should be on the order of a few tens of micro-Gauss (see figs. \\ref{090902B-epsilonB}, \\ref{090902B-epsilonBlate} and \\ref{090510-epsilonB}). For these three bursts, at least, no magnetic dynamo is needed to operate behind the shock front to amplify the magnetic field. The data for the short burst (GRB 090510) are consistent with the medium in the vicinity of the burst (within $\\sim$1 pc) being uniform and with density less than 0.1 cm$^{-3}$; the data rules out a CSM where $n\\propto R^{-2}$. On the other hand, the data for one of the two long {\\it Fermi} bursts (GRB 080916C) prefers a wind like medium and the other (GRB 090902B) a uniform density medium; these conclusions are reached independently from late time afterglow data alone and from the early time high energy data projected to late time using the 4-D parameter space technique described in \\S3. It is also interesting to note that the power-law index of the energy distribution of injected electrons ($p$) in the shocked fluid, for all the three {\\it Fermi} bursts analyzed in this work, is $2.4$ to within the error of measurement, suggesting an agreement with the {\\it Fermi} acceleration of particles in highly relativistic shocks, e.g. Achterberg et al. (2001); a unique power-law index for electrons' distribution in highly relativistic shocks is not found in all simulations. The study of high energy emission close to the deceleration time of GRB jets is likely to shed light on the onset of collisionless shocks and particle acceleration process. It might seem surprising that we are able to fit all data (optical, X-ray, $\\lae10^2$MeV) for these three {\\it Fermi} bursts with just a few parameters for the external forward shock. This is in sharp contrast to Swift bursts which often display a variety of puzzling (poorly understood) features in their afterglow light curves. There are two reasons that these {\\it Fermi} bursts can be understood using a very simple model (external forward shock). (1) The data for the two long {\\it Fermi} bursts (080916C and 090902B) are not available during the first 1/2 day, and that is precisely the time frame when complicated features (plateau, etc., eg. Nousek et al. 2006, O'Brien et al. 2006) are seen in the X-ray afterglow light curves of Swift bursts (we note that the external forward shock model in its simplest form can't explain these features) --- however, the afterglow data at later times is almost invariably a smooth single (or double) power-law function that can be modeled by synchrotron emission from an external forward shock. (2) For very energetic GRBs --- the three bursts we have analyzed in this paper are among the brightest bursts in their class --- the progenitor star is likely to be completely destroyed leaving behind very little material to fall back onto the compact remnant at the center to fuel continued activity and give rise to complex features during the first few hours of the X-ray afterglow light curve (Kumar, Narayan \\& Johnson, 2008). To summarize, the GRB afterglow physics was simple in the decade preceding the launch of Swift, and then things became quite complicated, and now the {\\it Fermi} data might be helping to clear the fog and reveal the underlying simplicity once again." }, "0910/0910.2555.txt": { "abstract": "IceCube is an all-flavor, cubic kilometer neutrino telescope currently under construction in the deep glacial ice at the South Pole. Its embedded optical sensors detect Cherenkov light from charged particles produced in neutrino interactions in the ice. For several years IceCube has been detecting muon tracks from charged-current muon neutrino interactions. However, IceCube has yet to observe the electromagnetic or hadronic particle showers or ``cascades'' initiated by charged-current or neutral-current neutrino interactions. The first detection of such an event signature is expected to come from the known flux of atmospheric electron and muon neutrinos. A search for atmospheric neutrino-induced cascades was performed using 275.46 days of data from IceCube's 22-string configuration. Reconstruction and background rejection techniques were developed to reach, for the first time, a signal-to-background ratio $\\sim$1. Above a reconstructed energy of 5~TeV, 12 candidate events were observed in the full dataset. The signal expectation from the canonical Bartol atmospheric neutrino flux model is $5.63\\pm2.25$ events, while the expectation from the atmospheric neutrino flux as measured by IceCube's predecessor array AMANDA is $7.48\\pm1.50$ events. Quoted errors include the uncertainty on the flux only. While a conclusive detection can not yet be claimed because of a lack of background Monte Carlo statistics, the evidence that we are at the level of background suppression needed to see atmospheric neutrino-induced cascades is strong. In addition, one extremely interesting candidate event of energy 133~TeV survives all cuts and shows an intriguing double pulse structure in its waveforms that may signal the ``double bang'' of a tau neutrino interaction. \\abstractsignature ", "introduction": " ", "conclusions": "" }, "0910/0910.3111.txt": { "abstract": "% context heading (optional) {Methyl cyanide is an important trace molecule in star-forming regions. It is one of the more common molecules used to derive kinetic temperatures in such sources.} % aims heading (mandatory) {As preparatory work for Herschel, SOFIA, and in particular ALMA we want to improve the rest frequencies of the main as well as minor isotopologs of methyl cyanide.} % methods heading (mandatory) {The laboratory rotational spectrum of methyl cyanide in natural isotopic composition has been recorded up to 1.63~THz.} % results heading (mandatory) {Transitions with good signal-to-noise ratio could be identified for CH$_3$CN, $^{13}$CH$_3$CN, CH$_3^{13}$CN, CH$_3$C$^{15}$N, CH$_2$DCN, and $^{13}$CH$_3^{13}$CN in their ground vibrational states up to about 1.2~THz. The main isotopic species could be identified even in the highest frequency spectral recordings around 1.6~THz. The highest $J'$ quantum numbers included in the fit are 64 for $^{13}$CH$_3^{13}$CN and 89 for the main isotopic species. Greatly improved spectroscopic parameters have been obtained by fitting the present data together with previously reported transition frequencies.} % conclusions heading (optional) {The present data will be helpful to identify isotopologs of methyl cyanide in the higher frequency bands of instruments such as the recently launched Herschel satellite, the upcoming airplane mission SOFIA or the radio telescope array ALMA.} ", "introduction": "\\label{intro} Methyl cyanide, CH$_3$CN, also known as acetonitrile, or cyanomethane, was among the early molecules to be detected by radio astronomical means. \\citet{det-MeCN} detected it almost 40 years ago toward the massive star-forming regions Sagittarius A and B close to the Galactic center. It has been detected in dark clouds such as TMC-1 \\citep{MeCN-TMC-1}, around the low-mass protostar IRAS 16293$-$2422 \\citep{MeCN_IRAS16293}, in the circumstellar shell of the famous carbon-rich star IRC~+10216 \\citep{MeCN_etc_IRC+10216}, in comets, such as Kohoutek \\citep{MeCN-Kohoutek}, and in external galaxies \\citep{MeCN_MeCCH_extragal}. As a strongly prolate symmetric top, meaning $A \\gg B$, transitions with different $K$ but the same $J$ occur in fairly narrow frequency regions, but sample rather different energies. Therefore, CH$_3$CN is quite commonly used to derive the kinetic temperatures of dense molecular clouds as already done in \\citet{det-MeCN} or in e.\\,g. \\citet{MeCN-T_kin}. In less dense clouds the large dipole moment of 3.922~D \\citep{MeCN-dipole} may prevent thermal equilibration. In that case, the isoelectronic methyl acetylene, or propyne, CH$_3$CCH, is often used instead. The $^{12}$C/$^{13}$C ratio in the vicinity of the Galactic center is approximately 20 \\citep{isotopic_abundances,13C-VyCN_2008}. Thus, it is not surprising that the two isotopomers containing $^{13}$C were detected fairly soon after the detection of the main isotopolog. CH$_3 ^{13}$CN \\citep{MeCN-T_kin} was detected first probably by accident because the isotopic substitution at the central C-atom does not change the spectroscopic parameters much with respect to the main species. \\citet{Orion-survey_13CH3CN} detected lines of both isotopomers with one $^{13}$C in a line survey of Orion. More recently, CH$_2$DCN has been detected in two hot core sources \\citep{det-CH2DCN}, and some lines of CH$_3$C$^{15}$N were detected in a line survey of Sagittarius~B2 \\citep{Sgr-B2_Nummelin}. A plethora of high-resolution spectroscopic investigations have been performed on methyl cyanide. It was actually among the first molecules to be studied by microwave spectroscopy by \\citet{MeCN_1st-MW}. \\citet{MeCN-LD} reviewed the investigations for the main isotopic species and performed high resolution, high accuracy measurements up to 1.2~THz. \\citet{MeCN_rot-isos_1996} analogously presented data up to 607~GHz for the singly substituted $^{13}$C isotopomers as well as for the species with $^{15}$N. \\citet{CH2DCN-rot} built on the very small data set for CH$_2$DCN by measuring an isotopically enriched sample up to 471~GHz. Finally, there has been only one report on the isotopolog with two $^{13}$C which was restricted to frequencies up to 72~GHz \\citep{MeCN-2x13}. We have measured rotational spectra of methyl cyanide in natural isotopic composition both in wide frequency windows as well as selected lines up to 1.63~THz to provide new or updated catalog entries for various isotopologs of methyl cyanide as well as for excited vibrational states. With rotational temperatures of CH$_3$CN in hot cores reaching at least 150$-$200~K \\citep{Orion-survey_13CH3CN,det-PrCN_EtFo}, these data will be of great relevance for the Atacama Large Millimeter Array (ALMA). The main isotopolog and possibly even the $^{13}$C isotopomers or the $\\varv _8 = 1$ excited vibrational state will be seen with the HIFI instrument (Heterodyne Instrument for the Far-Infrared) on board of the recently launched Herschel satellite or with SOFIA (Stratospheric Observatory For Infrared Astronomy). In the present article we provide the ground state rotational data obtained for six different isotopic species: CH$_3$CN, $^{13}$CH$_3$CN, CH$_3^{13}$CN, CH$_3$C$^{15}$N, CH$_2$DCN, and $^{13}$CH$_3^{13}$CN; as usual, unlabeled atoms refer to $^{12}$C and $^{14}$N. Analyses of selected excited vibrational states are currently under way. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% ", "conclusions": "\\label{conclusion} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% Rotational transitions for six astrophysically and astrochemically important isotopologs of methyl cyanide in their ground vibrational states have been recorded and the existing data sets have been extended greatly in most cases. The data should permit reliable predictions to above 2~THz in all instances, sufficient not only for Herschel or ALMA, but very likely also for such missions as SOFIA, CCAT, or others. Transitions of the even rarer isotopomers containing $^{15}$N and one $^{13}$C may be observable also, but chance of significant overlap are even higher and the chances to observe these species in space seem vanishingly low. Predictions of the rotational spectra of the isotopologs studied in the course of the present investigation will be available in the catalog section\\footnote{website: http://www.astro.uni-koeln.de/cdms/entries/, see also http://www.astro.uni-koeln.de/cdms/catalog/} of the Cologne Database for Molecular Spectroscopy\\footnote{website: http://www.astro.uni-koeln.de/cdms/} \\citep{CDMS_1,CDMS_2}. The complete line, parameter and fit files will be deposited in the Spectroscopy Data section of the CDMS. Updated or new JPL catalog \\citep{JPL-catalog} entries\\footnote{website: http://spec.jpl.nasa.gov/ftp/pub/catalog/catdir.html, see also http://spec.jpl.nasa.gov/} will be available also. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%" }, "0910/0910.1987_arXiv.txt": { "abstract": "Combination of Fresnel Zone Plates (FZP) can make an excellent telescope for imaging in X-rays. We present here the results of our experiments with several pairs of tungsten made Fresnel Zone plates in presence of an X-ray source kept at a distance of about $45$ feet. The quasi-parallel beam allowed us to study sources placed on the axis as well as off the axis of the telescope. We present theoretical study of the fringe patterns produced by the zone plates in presence of a quasi-parallel source. We compare the patterns obtained from experiments with those obtained by our Monte-Carlo simulations. The images are also reconstructed by deconvolution from both the patterns. We compare the performance of such a telescope with other X-ray imaging devices used in space-astronomy. ", "introduction": "\\label{intro} Imaging in X-rays in space has always been a challenging task. X-ray films are generally very inefficient and in space it is inconvenient to take continuous imaging with X-ray films. To circumvent this, Coded Aperture Masks (CAM) (Mertz \\& Young, 1960; Ables, 1968; Dicke, 1968) are widely used. While a CAM has an advantage that it is a single element instrument, it has the disadvantage that the resolution is low, and is limited by the smallest mask element. Recently, grazing incidence focusing instruments are being developed specially in space missions such as NUSTAR and SIMBOL-X (Ramsay, 2006; Pereschi, 2008), where target resolutions of $20-40$ arc seconds are being contemplated. A relatively simpler concept is, however, present in the literature. It was realized (Mertz 1965) that high resolution can be achieved when X-ray shadows are cast with two perfectly aligned Fresnel zone plates and the Moir\\'e fringes could be discerned by suitable detectors. Furthermore, the resolution is independent of the energy of X-rays. Theoretically, the resolution could be arbitrarily high as it increases with the separation $D$ of the zone plates, apart from being proportional to the finest element in the plate. Since small-pixel detectors were unavailable even though this concept was well established, there was hardly any application. Desai and his collaborators (1993, 1998, 2000) showed renewed interests in this type of telescopes and actually carried out an experiment and showed that an image reconstruction was possible. However, so far, a detailed and quantitative exploration of imaging with single and multiple sources in laboratory environments has not been carried out with Fresnel Zone Plate telescopes. With the recent development of CMOS based X-ray detectors where each pixel could be as small as $50$ microns, it has become necessary to revisit the problem once more to actually study the feasibility of sending Zone Plate telescopes for space applications. Furthermore, the high resolution imaging devices could be used for medical purposes also. This is important since a high dose of X-ray is required when conventional X-ray films are used. In laboratory and medical uses, however, the beams are likely to be diverging. It is therefore necessary to study the feasibility of imaging with diverging beams as well. In this series of papers, we plan to explore the practical aspects of the zone plate telescopes, focusing on the effects of the divergence of the beam and achievable resolutions in diverging beams. We also study with various combinations of zone plates. Our results are also supported by Monte-Carlo simulations carried out with parallel and divergent beams. Later in this series we will describe the results of the test and evaluation of two zone plate telescopes for the RT-2 experiments on board the Russian satellite CORONAS-PHOTON to be launched shortly. The plan of our present paper is the following. In the next Section, we briefly describe the mathematical formalism behind the functioning of a Fresnel zone plate telescope and the properties of the Moir\\'e patterns falling on the detector. In Sections 3 and 4, we describe the setup of our experiment and the scheme of Monte-Carlo simulations. In Section 5, we present results obtained with a single on-axis or off-axis source with two different configurations of the zone plates and with two types of zone plates. Monte-Carlo simulations of these cases are also presented to support the experimental results. In Section 6, we present results with multiple sources and present simulation results. Finally, in Section 7, we make concluding remarks. In Paper II, of this series we will make a comparative study of Monte-Carlo simulations for single and multiple sources at various distances as well as for parallel beam, which is not achievable in the laboratories. In Paper III (Nandi et al., 2008), we will discuss the test and evaluations of the RT-2 payload to be launched aboard CORONAS-PHOTON satellite. ", "conclusions": "Zone plates have long been conceived to generate high resolution achromatic X-ray images. Its resolution is superior compared to that of the Coded Aperture Masks (CAMs) because of its finer zones. In the present paper, we concentrated on the results of an experimental set-up where the X-ray beam is almost parallel. The collimator is $45$ feet long and a zone plate of $15$ mm radius makes an angle of less then $4$ arc minutes at the source. We derived the nature of the fringes theoretically and showed that a point source will produce concentric circles on the detector plane. Through experiments and through Monte-Carlo simulations, we have been able to present the fringe patterns and also reconstructed the corresponding images both for the on-axis and off-axis sources. Partly because of the diverging nature of the beam and partly due to the finite size of the detector pixels, the ideal theoretical resolution was not achieved. This could however be improved during further image processing. Resolution of a zone plate telescope could be made very high, since it is inversely proportional to the distance between the plates. From a practical point of view, in an astronomical satellite, a distance of a meter is possible which will yield (with smallest zones around $40-50$ micron as in our experiment) a ideal resolution of $8-10$ arc seconds. Another widely accepted type of X-ray imaging instrument having a very high resolution is the focusing grazing incidence type telescopes which are being contemplated in future space missions such as NUSTAR and SIMBOL-X. While these instruments target angular resolutions of $40$ and $20$ arc seconds respectively, to achieve these at high energies, formation flight architectures with focal lengths of $10$m and $20$m are required. Moreover, grazing incidence telescopes are expensive to build and the resolution is energy dependent, even when we assume the pointing is very accurate (any grazing type instrument is highly sensitive to alignment). In passing, we may remark that at a $20$m separation, our Zone plate telescope will definitely achieve an achromatic resolution of less than an arc second. One disadvantage of the Zone plate telescopes is that if the size of the finest zone and the distance of separation of the plates are kept fixed, the resolution will not increase even if the areas of the plates are made bigger. However, a larger size will gather more photons and will increase the sensitivity of the instrument. An achievable zone plate size could be very big (say one meter diameter, made by low $Z$ metal quoted with high Z metal, such as lead or gold). Even with CZT detectors with 2.5mm pixel, reasonably high resolution would be achieved. In order to check the efficiency of a zone plate telescope, we note that with our set up, simulations show (Paper-II) that a flux of about $10-12$ photons per cm$^2$ is required to have the source resolved about the background. With an increase in area this number does not change! For the crab nebula, the photon flux in $20-100$ keV range is about $0.12$ counts/sec/cm$^2$. Thus we need to integrate about $100$ seconds for imaging purpose. However, for a transient source, our instrument would be more effective since we could perhaps see change in size of the X-ray emitting region (such as glowing of the whole accretion disk when the accretion rate suddenly rises). With a larger field of view (as opposed to the focusing instruments), zone plate telescopes are ideally suited for imaging transient sources at a high energy. Of course, to eliminate high energy cosmic rays which will increase the background we require the usual protective shielding on our instrument. In the next paper (Palit et al. 2008), of this series we shall concentrate on the image resolution, and how it depends on the source distance and the detector details. We shall also demonstrate the capability of a complete zone plate telescope which will constitute four sets of zone plates. In Paper III, we shall concentrate on the setups used in the RT-2/CZT-CMOS payload aboard the Russian solar satellite CORONAS-PHOTON (Nandi et al. 2008)." }, "0910/0910.1349_arXiv.txt": { "abstract": "We present milliarcsecond-resolution radio very long baseline interferometry (VLBI) observations of the ultracool dwarfs TVLM\\,513--46546 (M8.5) and 2MASS J00361617+1821104 (L3.5) in an attempt to detect sub-stellar companions via direct imaging or reflex motion. Both objects are known radio emitters with strong evidence for periodic emission on timescales of about 2 and 3 hours, respectively. Using the inner seven VLBA antennas, we detect unresolved emission from TVLM\\,513--46546 on a scale of 2.5 mas ($\\sim 50$ stellar radii), leading to a direct limit on the radio emission brightness temperature of $T_B\\gtrsim 4\\times 10^5$ K. However, with the higher spatial resolution afforded by the full VLBA we find that the source appears to be marginally and asymmetrically resolved at a low S/N ratio, possibly indicating that TVLM\\,513--46546 is a binary with a projected separation of $\\sim 1$ mas ($\\sim 20$ stellar radii). Using the 7-hr baseline of our observation we find no astrometric shift in the position of TVLM\\,513--46546, with a $3\\sigma$ limit of about 0.6 mas. This is about 3 times larger than expected for an equal mass companion with a few-hour orbital period. Future monitoring of its position on a range of timescales will provide the required astrometric sensitivity to detect a planetary companion with a mass of $\\sim 10$ M$_J$ in a $\\gtrsim 15$ d ($\\gtrsim 0.06$ AU) orbit, or with a mass of $\\sim 2$ M$_J$ in an orbit of $\\gtrsim 0.5$ yr ($\\gtrsim 0.3$ AU). ", "introduction": "In recent years unexpectedly strong radio emission has been detected from very low mass stars and brown dwarfs (hereafter, ultracool dwarfs), revealing that these objects are capable of generating and dissipating kG-strength magnetic fields (e.g., \\citealt{ber01,ber06,hal08}). The radio luminosity remains nearly uniform from early-M to mid-L dwarfs, even though other magnetic activity indicators (H$\\alpha$ and X-rays) decline by about two orders of magnitude over the same spectral type range (e.g., \\citealt{whw+04,ber06}). Thus, radio observations provide a particularly powerful probe of the magnetic field properties. The radio emission from several ultracool dwarfs also exhibits a periodicity that matches the stellar rotation velocity ($v\\,{\\rm sin} i$), thereby pointing to a large-scale, axisymmetric, and stable magnetic field configuration on timescales of hours to years \\citep{ber05,hal06,hal07,ber09}. These conclusions are also supported by observations of Zeeman broadening in FeH molecular lines \\citep{rei07} and time-resolved optical spectropolarimetry \\citep{dmp08,mdp08}. Since ultracool dwarfs are fully convective, the solar-type $\\alpha\\Omega$ dynamo, which operates at the shearing interface between the radiative and convective zones \\citep{par55}, cannot be responsible for generating and amplifying the inferred fields. However, recent numerical simulations suggest that large scale axisymmetric fields can still be generated in fully convective objects, at least for conditions that roughly correspond to mid-M dwarfs ($M\\sim 0.3$ M$_\\odot$; \\citealt{bro08}). Simulations that correspond to the conditions in ultracool dwarfs are challenging and have not been investigated so far. As a result, observational constraints on the scale, geometry, and origin of the fields are essential. The spectroscopic Zeeman techniques are currently of limited utility for the dim and generally rapidly-rotating ultracool dwarfs, since they require very high signal-to-noise ratios and have reduced sensitivity at high rotation velocities due to line broadening \\citep{rei07}. Detailed radio observations are thus crucial. Here we present the first very long baseline interferometry (VLBI) observations of ultracool dwarfs, aimed at providing further constraints on their magnetic field properties. In particular, these observations have sufficient angular resolution to detect astrometric shifts of a few tenths of a milliarcsecond that may be exerted by a sub-stellar companion with a period of a few hours. They also allow us to directly image a radio-emitting companion down to milliarcsecond scales. The presence of such putative companions may enhance the magnetic activity through direct or tidal interaction. We detect the well-studied M8.5 dwarf TVLM\\,513-46546 with the VLBA, the first such detection for any ultracool dwarf. The detectability of this object with the VLBA suggests that long-term astrometric monitoring could reveal the presence of a planetary-mass companion on a $\\gtrsim 15$ d orbit. Such a detection would also usher in a new era of extrasolar planet studies through radio astrometry. ", "conclusions": "Using high angular resolution radio VLBI observations, we clearly detect the ultracool dwarf TVLM513--46 on an unprecedented small scale of about 2.5 mas or about 50 stellar radii. The source is unresolved on this scale, but it may be marginally resolved on a $\\sim 1$ mas scale when using the longest VLBA baselines, although the S/N ratio is low. Given the corresponding physical scale of about 20 stellar radii, this may point to a binary system with two radio-emitting sources in a $\\gtrsim 1$ d orbit. Beyond the possibility of a spatially resolved binary, a search for an astrometric shift in the position of the source due to the gravitational influence of a close companion yields a $3\\sigma$ limit of $\\lesssim 0.6$ mas, about 3 times larger than the expected shift for an equal-mass companion in a few-hour orbit. Since we lost the phased-VLA and GBT data, we expect that future VLBI observations will provide the requisite astrometric accuracy to search for an equal-mass companion in such a tight orbit. These future observations will also allow us to determine whether the marginally-resolved emission from TVLM513--46 is indeed due to an equal mass companion with a projected separation of about 20 stellar radii. If this is indeed the case, the expected maximum reflex motion is about 1 mas, easily detectable with VLBA+GBT observations that span several days. Equally important, the fact that TVLM513--46 is detected with the VLBA opens the possibility to astrometrically search for a planetary-mass companion with a $\\gtrsim 10$ d orbit. Monitoring of TVLM513--46 with the VLBA or VLBA+GBT on timescales of days to months to years will allow us to probe the existence of companions with masses of $\\sim 1-10$ M$_J$ and with orbits of $\\sim 15$ d to $\\sim 1$ yr. These observations may lead to the first detection of an extrasolar planet via radio astrometry. Thus, the use of VLBI observations opens up a wide companion parameter space for TVLM513--46, with detection via direct imaging for a roughly equal-mass radio-emitting companion down to a scale of $\\sim 20$ stellar radii, and detection via reflex motion on scales as small as a few stellar radii for an equal mass companion, and scales of $\\gtrsim 0.1$ AU for a planetary mass companion. This first VLBA detection also paves the way for similar observations of known and future radio-emitting ultracool dwarfs in search of companions." }, "0910/0910.2383_arXiv.txt": { "abstract": "Reported observations in H$\\alpha$, Ca II H and K or or other chromospheric lines of coronal rain trace back to the days of the Skylab mission. Offering a high contrast in intensity with respect to the background (either bright in emission if observed at the limb, or dark in absorption if observed on disk) these cool blobs are often observed falling down from high coronal heights above active regions. A physical explanation for this spectacular phenomenon has been put forward thanks to numerical simulations of loops with footpoint concentrated heating, a heating scenario in which cool condensations naturally form in the corona. This effect has been termed ``catastrophic cooling'' and is the predominant explanation for coronal rain. In this work we further investigate the link between this phenomenon and the heating mechanisms acting in the corona. We start by analyzing observations of coronal rain at the limb in the Ca II H line performed by the \\textit{Hinode} satellite. We then compare the observations with 1.5-dimensional MHD simulations of loops being heated by small-scale discrete events concentrated towards the footpoints (that could come, for instance, from magnetic reconnection events), and by Alfv\\'en waves generated at the photosphere. It is found that if a loop is heated predominantly from Alfv\\'en waves coronal rain is inhibited due to the characteristic uniform heating they produce. Hence coronal rain may not only point to the spatial distribution of the heating in coronal loops but also to the agent of the heating itself. We thus propose coronal rain as a marker for coronal heating mechanisms. ", "introduction": "Coronal loops are dynamical entities which exhibit heating and cooling processes constantly. Most of them are actually far from being in hydrostatic equilibrium \\citep{Aschwanden_2001ApJ...550.1036A}. The dynamical nature of loops can manifest in observations in a variety of forms, among which propagating intensity variations is a common one. Indeed, intensity variations traveling along coronal loops have been frequently observed in many different wavelengths, in hot coronal EUV lines as well as chromospheric cool lines. The agents producing these intensity variations can be either propagating waves or flows along coronal loops, and observers can have a hard time differentiating them despite their very different physical nature. Coronal rain is an example of intensity variations caused by flows. This spectacular phenomenon corresponds to cool plasma condensations (with chromospheric temperatures) in a hot environment (coronal temperatures) falling down along coronal loops, hence the very appropriate name. Coronal rain seems to be a common phenomenon above active regions, where loops are dense. Due to the low temperatures, the condensations appear as bright emission profiles in chromospheric lines such as H$\\alpha$ or Ca II H and K when observed at the limb \\citep[][Fig. \\ref{fig1}]{DeGroof_2004AA...415.1141D} or as dark absorption profiles when observed on disk. If observed in EUV wavelengths these propagating blobs appear as dark features \\citep{Schrijver_2001SoPh..198..325S}. \\citet{DeGroof_2004AA...415.1141D, DeGroof05} report simultaneous observations in EIT 304 \\AA, H$\\alpha$ images from Big Bear Solar Observatory, and in 171 \\AA~ (\\textit{TRACE}) of intensity variations propagating along a coronal loop. Following a detailed analysis they put forward a series of points through which a differentiation between waves and flows (cool plasma condensations, coronal rain) can be done. Propagating waves frequently reported in EIT/\\textit{SOHO} 195 \\AA~ and \\textit{TRACE} 171 \\AA~ appear as low min-to-max intensity variations as compared to coronal rain. Most observed waves correspond to sound modes propagating at constant speed corresponding to the local sound speed of the corona, $\\sim$ 150 km s$^{-1}$. On the other hand, cool plasma condensations are seen to accelerate while falling along coronal loops to velocities above 100 km s$^{-1}$. Furthermore, waves have only been seen to propagate upwards, within the first 20 Mm of loops, after which they are usually damped, while cool condensations have only been seen to fall down along the loops from coronal heights. The falling speeds of the condensations have been reported to be lower than free fall speeds, due probably to the gas pressure along the loop, which increases considerably below the transition region. In some cases continuous flows from one footpoint to the other are also observed. \\citet{Antiochos_1999ApJ...512..985A} propose a common physical mechanism for coronal rain and prominence formation. Ranges for temperatures and densities seem to be shared by both mechanisms, as well as the location of formation, high in the corona. While a prominence forms and is supported by the magnetic field for a time scale of days, coronal rain occurs on a time scale of minutes. It was shown that prominences may not need to rely on the geometry of the magnetic field to form (presence of ``dips''), contrary to the general belief. They showed that a coronal loop whose heating is concentrated towards the footpoints is subject to a thermal instability in the corona, which dramatically cools down to chromospheric temperatures in time scales of minutes once the density and temperatures have reached critical values. This phenomenon was termed ``catastrophic cooling'' and has so far gained acceptance as possible explanation for coronal rain. This mechanism occurs in a cyclic manner, matching reported periodicity of coronal rain observations \\citep{Schrijver_2001SoPh..198..325S}. This periodicity depends on geometrical aspects such as the loop length and the heating scale height, and can be obtained even with a time independent heating \\citep{Muller_2003AA...411..605M}. As the heating is concentrated towards the footpoints a larger mass flux is input into the corona as compared to uniform heating. The density of the corona increases in time while the energy flux is kept constant due to the constant heating input at the footpoints. This causes the loop to reach a maximum temperature after which it starts decreasing due to the decreasing of the heating per unit mass. The lower temperatures and higher densities increase the radiative losses which further cool the corona. Eventually, in a time scale of hours, the loop reaches a critical state and a thermal instability sets in. Then, in a time scale of minutes the temperature and density respectively decrease and increase dramatically to chromospheric values. The cool condensation subsequently falls down subject to gravity and plasma pressure. The loop subsequently is depleted and reheats rapidly due to the low density. The cycle then repeats. We will refer to these cycles as ``limit cycles\", as termed by \\citep{Muller_2003AA...411..605M}. Observations seem to indicate that Coronal rain is a fairly common phenomenon of active regions, where loops are hot and dense. The coronal heating mechanism being a great unknown in solar physics, it is interesting to think about its underlying relationship to coronal rain. At a first glance coronal rain may appear as a random failure of the coronal heating mechanism to heat the loops in active regions. However, as previously stated, this phenomenon seems to act in the corona in a cyclic manner. Furthermore simulations have pointed out to the necessity of specific spatially dependent heating in order to allow the catastrophic cooling leading to coronal rain \\citep{Antiochos_1999ApJ...512..985A, Muller_2003AA...411..605M}. This cooling phenomenon seems then to be deeply linked to the unknown coronal heating mechanism. Since it is a fairly easily observable phenomenon due to the large velocities and density variation (hence clear Doppler shifted emission or absorption profiles), it may act as a marker of the operating heating mechanism in the loop. It has been shown for instance that footpoint concentrated heating may lead to catastrophic cooling if the heating scale height is sufficiently concentrated towards the footpoints \\citep{Muller_2003AA...411..605M}. Uniform heating on the other hand fails to reproduce the phenomenon since the heating rate per unit mass needs to decrease in time locally in the corona in order to allow the thermal instability to set in. It was further shown that catastrophic cooling does not need time dependent heating. In other words, it can happen even with a constant heating function. The footpoint concentrated heating function to which simulations point to matches the observational evidence of coronal loops above active regions for being mainly heated at their footpoints \\citep{Aschwanden_2001ApJ...560.1035A}, which sets most loops out of hydrostatic equilibrium. Further evidence of this fact has been found by \\citet{Hara_2008ApJ...678L..67H} using the Hinode/EIS instrument, which shows that active region loops exhibit upflow motions and enhanced nonthermal velocities. Possible unresolved high-speed upflows were also found, fitting in the footpoint concentrated heating scenario. Many heating mechanisms have been proposed as candidates for heating the solar corona up to the observed few million degree temperatures. In this context a large emphasis has been put on the search for Alfv\\'en waves in the solar corona. Theoretically, they can be easily generated in the photosphere by the constant turbulent convective motions, which inputs large amounts of energy into the waves \\citep{Muller_1994AA...283..232M, Choudhuri_1993SoPh..143...49C}. Having magnetic tension as its restoring force the Alfv\\'en waves can travel less affected by the large transition region gradients with respect to other modes. Also, when traveling along thin magnetic flux tubes they are cut-off free since they are not coupled to gravity (Musielak et al. 2007)\\footnote{\\citet{Verth_etal_aap_09} have pointed out however that the assertion made by \\citet{Musielak_2007ApJ...659..650M} is valid only when the temperature in the flux tube does not differ from that of the external plasma. When this is not the case a cut-off frequency is introduced.}. Alfv\\'en waves generated in the photosphere are thus able to carry sufficient energy into the corona to compensate the losses due to radiation and conduction, and, if given a suitable dissipation mechanism, heat the plasma to the high million degree coronal temperatures \\citep{Uchida_1974SoPh...35..451U, Wentzel_1974SoPh...39..129W, Hollweg_1982SoPh...75...35H, Poedts_1989SoPh..123...83P, Ruderman_1997AA...320..305R, Kudoh_1999ApJ...514..493K, Antolin_2009arXiv0910.0962v1} and power the solar wind \\citep{Suzuki_2006JGRA..11106101S, Cranmer_2007ApJS..171..520C}. The main problem faced by Alfv\\'en wave heating is actually to find a suitable dissipation mechanism. Being an incompressible wave it must rely on a mechanism by which to convert the magnetic energy into heat. Several dissipation mechanisms have been proposed, such as parametric decay \\citep{Goldstein_1978ApJ...219..700G, Terasawa_1986JGR....91.4171T}, mode conversion \\citep{Hollweg_1982SoPh...75...35H, Kudoh_1999ApJ...514..493K, Moriyasu_2004ApJ...601L.107M}, phase mixing \\citep{Heyvaerts_1983AA...117..220H, Ofman_2002ApJ...576L.153O}, or resonant absorption \\citep{Ionson_1978ApJ...226..650I, Hollweg_1984ApJ...277..392H, Poedts_1989SoPh..123...83P, Erdelyi_1995AA...294..575E}. The main difficulty lies in dissipating sufficient amounts of energy in the correct time and space scales. For more discussion regarding this issue the reader can consult for instance \\citet{Klimchuk_2006SoPh..234...41K}, \\citet{Erdelyi_2007AN....328..726E} and \\citet{Aschwanden_2004psci.book.....A}. Works considering Alfv\\'en wave heating as coronal heating mechanism have shown that the obtained coronae are uniformly heated \\citep{Moriyasu_2004ApJ...601L.107M, Antolin_2008ApJ...688..669A, Antolin_2009arXiv0910.0962v1, Suzuki_2006JGRA..11106101S}. In these works the heating issues from shocks of longitudinal modes (mainly slow modes) from mode conversion of the Alfv\\'en waves due to the density fluctuation, wave-to-wave interaction and deformation of the wave shape during propagation. The coronal loops issuing from Alfv\\'en wave heating are found to satisfy quite well the RTV scaling law \\citep{Rosner_1978ApJ...220..643R} which quantifies the heating uniformity in the loops. This result would point then towards an inhibition of coronal rain if Alfv\\'en wave heating is predominant in the loop. It is this idea that is addressed in this work. Another promising coronal heating candidate mechanism is the nanoflare reconnection heating model. The nanoflare reconnection process was first suggested by \\citet{Parker_1988ApJ...330..474P}, who considered a magnetic flux tube as being composed by a myriad of magnetic field lines braided into each other by continuous footpoint shuffling. Many current sheets in the magnetic flux tube would be created randomly along the tube that would lead to many magnetic reconnection events, releasing energy impulsively and sporadically in small quantities of the order of $10^{24}$ erg or less (nanoflares). Parker's original idea was a nanoflare reconnection heating acting uniformly in the corona but it was later proposed to be concentrated towards the footpoints of loops, where the magnetic canopy lies and magnetic field lines may entangle \\citep{Klimchuk_2006SoPh..234...41K, Aschwanden_2001ApJ...560.1035A}. In the reconnection scenario waves are also expected to be generated, and the energy imparted into Alfv\\'en waves is a matter of debate. The imparted energy may well depend on the location in the atmosphere of the reconnection event. \\citet{Parker_1991ApJ...372..719P} suggested a model in which 20~\\% of the energy released by reconnection events in the solar corona is transfered as a form of Alfv\\'en wave.\\citet{Yokoyama_1998ESASP.421..215Y} studied the problem simulating reconnection in the corona, and found that less than 10~\\% of the total released energy goes into Alfv\\'en waves. This result is similar to the 2-D simulation results of photospheric reconnection by \\citet{Takeuchi_2001ApJ...546L..73T}, in which it is shown that the energy flux carried by the slow magnetoacoustic waves is one order of magnitude higher that the energy flux carried by Alfv\\'en waves. On the other hand, recent simulations by Kigure et al. (private communication) show that the fraction of Alfv\\'en wave energy flux in the total released magnetic energy during reconnection in low $\\beta$ plasmas may be significant (more than 50~\\%). Since the observed ubiquitous intensity bursts (nanoflares) are thought to play an important role in the heating of the corona \\citep{Hudson_1991SoPh..133..357H} and since they are generally assumed to be a signature of magnetic reconnection it is then crucial to determine the energy going into the Alfv\\'en waves during the reconnection process. Moreover, \\citet{Moriyasu_2004ApJ...601L.107M} has shown that the observed spiky intensity profiles due to impulsive releases of energy may actually be a signature of Alfv\\'en waves. It was found that due to nonlinear effects Alfv\\'en waves can convert into slow and fast magnetoacoustic modes which then steepen into shocks and heat the plasma to coronal temperatures balancing losses due to thermal conduction and radiation. The shock heating due to the conversion of Alfv\\'en waves was found to be episodic, impulsive and uniformly distributed throughout the corona, producing an X-ray intensity profile that matches observations. Hence, \\citet{Moriyasu_2004ApJ...601L.107M} proposed that the observed nanoflares may not be directly related to reconnection but rather to Alfv\\'en waves. Differentiating Alfv\\'en wave heating from nanoflare reconnection heating during observations is one of the main tasks needed in order to solve the coronal heating problem. Following the work of \\citet{Moriyasu_2004ApJ...601L.107M}, \\citet{Antolin_2008ApJ...688..669A} have compared both heating mechanisms by studying the hydrodynamic response of a loop subject to both kinds of heating. It was found that Alfv\\'en waves lead to a dynamic, uniformly heated corona with steep power law indexes (issuing from statistics of heating events) while nanoflare-reconnection heating leads to lower dynamics (besides the times when catastrophic cooling takes place in the case of footpoint concentrated heating) and shallow power laws. It was further found that footpoint nanoflare heating (namely, nanoflare reconnection heating concentrated towards the footpoints of loops) leads to hot upflows (as observed in the Fe\\,XV 284.16~\\AA\\ line) due to the plasma being heated rapidly towards the footpoints before being ejected into the corona, while Alfv\\'en wave heating leads to hot downflows due to the plasma achieving the maximum temperatures in the corona (rather than at the footpoints) and being carried back to the footpoints by the strong shocks generated there \\citet{Antolin_2009arXiv0903.1766A}. In this work we propose coronal rain as another observational signature through which both heating mechanisms can be distinguished. We start first by reporting limb observations of coronal rain from \\textit{Hinode}/SOT in the Ca\\,II H line. The velocities and shapes of the falling condensations are analyzed. With a 1.5-dimensional code we then proceed to model a coronal loop being subject to a heating mechanism that is concentrated towards the footpoints, such as the nanoflare reconnection heating model \\citep[as proposed by][]{Aschwanden_2001ApJ...560.1035A, Klimchuk_2006SoPh..234...41K}. In the case considered catastrophic cooling happens 3 times and we select the first event to analyze and compare with our observations. Next we proceed by generating Alfv\\'en waves at the photosphere and gradually increase the amplitude. As the heating from the waves becomes non-negligible with respect to the heating flux from the nanoflare heating events catastrophic cooling is inhibited and the loop reaches a thermal equilibrium state. This result has important consequences since it points to the conclusion that Alfv\\'en waves are a non predominant heating mechanism in active region loops. The work is organized as follows. In \\S\\ref{observations} we report observations of coronal rain in the Ca II H line performed with \\textit{Hinode}/SOT. In \\S\\ref{model} we introduce the 1.5-dimensional MHD model in which our loop is based on and discuss the heating models of the loop. In \\S\\ref{simulation} we present the results of footpoint heating and analyze a typical case of catastrophic cooling. The effect of Alfv\\'en waves on the thermal stability of the loop is also studied. In \\S\\ref{discussion} we discuss the results in the context of coronal heating and conclude the work. ", "conclusions": "\\label{discussion} Observations in chromospheric lines such as H$\\alpha$ or Ca II H and K seem to show that coronal rain is a phenomenon exclusively of active regions \\citep{Schrijver_2001SoPh..198..325S, DeGroof_2004AA...415.1141D}, where loops are dense and heating appears to be concentrated towards the footpoints \\citep{Aschwanden_2001ApJ...550.1036A, Aschwanden_2001ApJ...560.1035A, Hara_2008ApJ...678L..67H}. In this work we have reported observations with the Hinode/SOT telescope in the Ca II H line of coronal rain occurring over an active region, and compared with results of simulations of loops that undergo catastrophic cooling. We have found that catastrophic cooling of coronal loops satisfactorily reproduces the main observed features of coronal rain. The loops are preferentially heated towards the footpoints and are subject to cycles \\citep[``limit cycles'', as termed by][]{Muller_2003AA...411..605M} in which they rapidly cool down and then reheat, they get dense and then deplete. The constant heating input at the footpoints of the loops produces coronae out of hydrostatic equilibrium. The coronal density increases in time causing a gradual decrease of the temperature and increase of the radiative losses. When the temperatures are sufficiently low radiative losses are dramatically increased since the condensation becomes optically thick and radiates considerably more. Catastrophic cooling sets in, either locally in the corona or globally, case in which the entire corona is cooled down to chromospheric values. Dense condensations of cool plasma form at coronal heights, which subsequently fall down by gravity. The loop is then depleted and the cycle starts again. The characteristics of our condensation match well the main characteristics of the coronal rain from the reported observations. The velocities are in the same range as would be expected from coronal rain forming in roughly the same locations of loops having essentially the same length. We obtain an elongation of the condensation in the simulation as it gets denser and the effective gravity increases along its way down the loop. This elongation is observed as well in the present Hinode observations and has previously also been reported \\cite{Schrijver_2001SoPh..198..325S}. The temperatures and densities of the resulting condensation have chromospheric values, which suggest that they would emit radiation in lines such as H$\\alpha$ or Ca II H or K, as in the observations with Hinode/SOT or the Swedish Solar Telescope. In a future paper we will test this idea by synthesizing the H$\\alpha$ line profile based on self-consistent solution of the ionization rate equations coupled with the hydrodynamic equations. As suggested by the simulations coronal rain is subject not only to gravity but also to the pressure changes inside the loop. Catastrophic cooling is the abrupt loss of thermal equilibrium in the corona, through which large pressure changes along the loop can occur and which can drive strong flows and shocks along the loop. In the catastrophic cooling event from the simulation reported here an upward motion of the condensation is obtained due to the local changes in pressure from the flows and the shocks. This may be a possible explanation to the upward motion of coronal rain as reported from the observations with Hinode/SOT in the Ca II H line. If coronal rain is indeed the consequence of the catastrophic cooling mechanism we then must have a heating mechanism acting preferentially towards the footpoints in loops where coronal rain is observed, i.e. active region loops. We have seen that Alfv\\'en waves generated at the footpoints of loops produce uniform coronae and are thus unable to reproduce phenomena such as coronal rain. Furthermore, when Alfv\\'en waves are present in a loop and have enough energy flux to heat and maintain a hot corona the catastrophic cooling events are inhibited. The loop then converges to a uniform and steady state. Coronal loops in active regions show a recurrent occurrence of coronal rain. They are dynamical entities showing heating and cooling processes at all times. Our results indicate that having Alfv\\'en wave heating as the main heating mechanism in loops means the absence of coronal rain. These results indicate then that Alfv\\'en wave heating may not be the principal heating mechanism for coronal loops in active regions. \\citet{Hara_2008ApJ...678L..67H}, using the \\textit{Hinode}/EIS instrument, have found upflow motions and enhanced nonthermal velocities in the hot lines of Fe XIV 274 and Fe XV 284 in active region loops. Possible unresolved high-speed upflows were also found. In \\citet{Antolin_2008ApJ...688..669A} and \\citet{Antolin_2009arXiv0903.1766A} we found that footpoint or uniform heating coming from nanoflare reconnection exhibit hot upflows, thus fitting in the observational scenario of active regions, while Alfv\\'en wave heating was found to exhibit hot downflows, which may fit in the observational scenario of quiet Sun regions \\citep{Chae_1998ApJS..114..151C, Brosius_2007ApJ...656L..41B}, further supporting the present conclusions. In \\citet{Antolin_2009arXiv0910.0962v1} we have found that Alfv\\'en wave heating is effective only in thick loops (with area expansions between photosphere and corona higher than 600) and long loops (with lengths of the corona above 50 Mm), a scenario which may not fit in active regions, where loops exhibit low area expansions due to the high magnetic field filling factors in those regions. Hence Alfv\\'en waves may play an important role in the heating of Quiet Sun regions (rather than active regions), where loops are often long, expand more than in active regions, kG (or higher) bright points are ubiquitous and coronal rain appears to be absent." }, "0910/0910.1312_arXiv.txt": { "abstract": "We describe an extension of the hadronic SU(3) non-linear sigma model to include quarks. As a result, we obtain an effective model which interpolates between hadronic and quark degrees of freedom. The new parameters and the potential for the Polyakov loop (used as the order parameter for deconfinement) are calibrated in order to fit lattice QCD data and reproduce the QCD phase diagram. Finally, the equation of state provided by the model, combined with gravity through the inclusion of general relativity, is used to make predictions for neutron stars. \\vspace{1pc} ", "introduction": "Dense matter hadronic models, such as the ones used to describe neutron stars, work in a determined range of energies \\cite{Glendenning:1991ic,Weber:1989uq,Schaffner:1995th}. They can, in addition, fulfill some symmetries of the underlying theory (QCD) such as chiral symmetry \\cite{chiral2,Heide:1993yz,Carter:1995zi} and scale invariance \\cite{Bonanno:2008tt,Bonanno:2009fg} but they do not include deconfinement to quark matter. On the other hand, the broadly used bag-model \\cite{bag0} includes quark degrees of freedom but does not include chiral symmetry. Models such as the quark-NJL and quark sigma-models \\cite{Nambu:1961tp,Nambu:1961fr,Bub} include that symmetry, but do not directly incorporate hadronic degrees of freedom or quark confinement. Our goal is to construct an effective model that contains hadronic and quark degrees of freedom present at different densities/temperatures but that can also coexist in a mixed phase. This allows the deconfinement phase transition to be a steep first order as well as a smooth crossover and cases in between. The last two possibilities, despite being predicted by lattice QCD, cannot be achieved by the usual procedure that connects, by hand, two different models with separate equations of state for hadronic and quark matter at the chemical potential at which the pressure of the quark EOS exceeds the hadronic one \\cite{hybrid1}. In order to achieve this goal, we employ a single model for the hadronic and quark phases. Our extension of the hadronic SU(3) non-linear sigma model also includes quark degrees of freedom in a spirit similar to the PNJL model \\cite{PNJL}, in the sense that it is a non-linear sigma model that uses the Polyakov loop $\\Phi$ as the order parameter for the deconfinement. This is a quite natural idea, since the Polyakov loop is related to the Z(3) symmetry, that is spontaneously broken by the presence of quarks. In QCD the Polyakov loop $\\Phi$ is defined via $\\Phi=\\frac13$Tr$[\\exp{(i\\int d\\tau A_4)}]$, where $A_4=iA_0$ is the temporal component of the SU(3) gauge field. ", "conclusions": "In spite of the fact that our model is relatively simple, it is the only one able to take into account different degrees of freedom and consequently allow steep as well as smooth transitions between different phases. The model is in accordance with lattice QCD data and the phase diagram it reproduces is able to describe a broad variety of regimes, from compact stars to heavy ion collisions. Calculations along this line are in progress \\cite{soon}. We conclude that our model is suitable for the description of neutron stars, since it predicts masses and radii in agreement with observations. Although it does not predict stable stars containing pure quark matter, the model allows stars that contain a core of mixed phase that can extend up to approximately $2$ km of radius. Even in this case, the reduced maximum star mass is still higher than the most massive pulsars observed. A major advantage of our work compared to other studies of hybrid stars is that we can study in detail the way in which chiral symmetry is restored and the way deconfinement occurs at high temperature/density. Since the properties of the physical system, e.g. the density of particles in each phase, are directly connected to the Polyakov loop it is not surprising that we obtain different results in a combined description of the degrees of freedom compared to a simple matching of two separate equations of state." }, "0910/0910.3317_arXiv.txt": { "abstract": "\\noindent An analysis of the first set of low-redshift ($z$$<$$0.08$) Type~Ia supernovae monitored by the Carnegie Supernova Project between 2004 and 2006 is presented. The data consist of well-sampled, high-precision optical ($ugriBV$) and near-infrared (NIR; $YJHK_s$) light curves in a well-understood photometric system. Methods are described for deriving light-curve parameters, and for building template light curves which are used to fit Type~Ia supernova data in the $ugriBVYJH$ bands. The intrinsic colors at maximum light are calibrated using a subsample of supernovae assumed to have suffered little or no reddening, enabling color excesses to be estimated for the full sample. The optical--NIR color excesses allow the properties of the reddening law in the host galaxies to be studied. A low average value of the total-to-selective absorption coefficient, $R_V \\approx 1.7$, is derived when using the entire sample of supernovae. However, when the two highly reddened supernovae (SN~2005A and SN~2006X) in the sample are excluded, a value $R_V \\approx 3.2$ is obtained, similar to the standard value for the Galaxy. The red colors of these two events are well matched by a model where multiple scattering of photons by circumstellar dust steepens the effective extinction law. The absolute peak magnitudes of the supernovae are studied in all bands using a two-parameter linear fit to the decline rates and the colors at maximum light, or alternatively, the color excesses. In both cases, similar results are obtained with dispersions in absolute magnitude of 0.12--0.16~mag, depending on the specific filter-color combination. In contrast to the results obtained from the comparison of the color excesses, these fits of absolute magnitude give $R_V \\approx$ 1--2 when the dispersion is minimized, even when the two highly reddened supernovae are excluded. This discrepancy suggests that, beyond the ``normal'' interstellar reddening produced in the host galaxies, there is an intrinsic dispersion in the colors of Type~Ia supernovae which is correlated with luminosity but independent of the decline rate. Finally, a Hubble diagram for the best-observed subsample of supernovae is produced by combining the results of the fits of absolute magnitude versus decline rate and color excess for each filter. The resulting scatter of 0.12~mag appears to be limited by the peculiar velocities of the host galaxies as evidenced by the strong correlation between the distance-modulus residuals observed in the individual filters. The implication is that the actual precision of Type~Ia supernovae distances is 3--4\\%. ", "introduction": "\\label{sec:intro} The discovery in the late-1990s from measurements of distant Type~Ia supernovae (SNe~Ia) that the expansion of the Universe is currently accelerating \\citep{riess98,perlmutter99} has given rise to an exciting new era in cosmology. The observations appear to require a previously unrecognized form of energy density (``dark energy'') which today is the dominant constituent of the Universe. In recent years, observations of SNe~Ia to increasingly higher redshifts have confirmed this finding \\citep[e.g.,][]{riess07,astier06,wood-vasey07}. Measurements of the angular power spectrum of the cosmic microwave background radiation provide independent confirmation, favoring a flat geometry for the Universe with a dominant 73\\% of the energy density in the form of dark energy \\citep[e.g., see][]{spergel07}. Baryon acoustic oscillations \\citep{eisenstein05} and measurements of the ratio of X-ray-emitting gas to total mass in galaxy clusters \\citep{allen04,allen07} have also provided convincing confirmation of this picture. Effort is currently focused on determining the nature of dark energy by measuring an equation-of-state parameter of the form $w = P/(\\rho c^2$), and its time derivative $\\dot{w}$. Several ongoing and future projects aim at filling the Hubble diagram with high-redshift ($z>0.1$) SNe~Ia in order to measure $w$ and $\\dot{w}$ with sufficient precision to distinguish among the various proposed models of dark energy. However, these measurements require an improved reference set of SN~Ia observations at low redshift. Surprisingly, the currently available low-redshift datasets are neither sufficiently numerous nor homogeneous to adequately complement the high-redshift data. The Carnegie Supernova Project (CSP) is one of several ongoing efforts to improve the quality of the low-redshift data.\\footnote[9]{See \\citet{contreras09} for a summary of other groups that are producing large databases of low-redshift SN~Ia light curves.} Over the five-year period ending in May~2009, the CSP has obtained densely sampled optical {\\em and} near-infrared (NIR) light curves of $\\sim$100 SNe~Ia in a well-understood, homogeneous photometric system \\citep[][hereafter H06]{hamuy06}. The NIR was included specifically to address the unknown intrinsic colors and interstellar reddening to SNe~Ia. This work presents an analysis of the first release of SNe~Ia by the CSP. In \\S~\\ref{sec:lcs}, we present an analysis of the light curves, both optical and NIR, and introduce the methods used to derive parameters which serve to determine distances. In \\S~\\ref{sec:redd}, the non-trivial issue of measuring extinction is considered, and the nature of the reddening law in the host galaxies is studied. In \\S~\\ref{sec:absmag}, the properties of SNe~Ia as {\\em standardizable\\,} candles is examined by using Hubble-flow distances to fit the relationship between absolute peak magnitudes, decline rates, and colors or reddenings for different bands. Finally, in \\S~\\ref{sec:concl}, the implications of our findings are discussed. This paper is the second of three companion papers presenting first results from CSP observations of low- and high-redshift SNe~Ia. The low-redshift data used in the present analysis are described in detail in the first paper by \\citet[][hereafter C09]{contreras09}. In the third paper, NIR light curves of 35 high-redshift SNe~Ia observed with the Magellan Baade 6.5-m telescope are combined with the low-redshift data in order to construct the first rest-frame $i$-band Hubble diagram of SNe~Ia \\citep{freedman09}. ", "conclusions": "\\label{sec:concl} This paper presents a first analysis of the light curves of 34 SNe~Ia followed by the CSP and released by C09. The high photometric precision and dense time sampling of the observations, especially in the optical bands, has allowed an in-depth examination of the general properties of the light curves of SNe~Ia in the $ugriBVYJH$ bands. Subsamples of well-observed SNe --- 26 in the optical and 9 in the NIR --- were used to build a family of template light curves using a technique that allows one to interpolate among the template data, taking into account variations in the sampling, in a 3-D space parameterized by the epoch relative to maximum light, the magnitude relative to maximum, and the decline rate. The availability of observations covering a wide range of wavelengths (from $u$ to $K_s$) offers the opportunity to make further progress on the difficult issue of correcting SN luminosities for extinction outside the Galaxy. However, depending on the approach taken, two quite different results are found. In the first case, a subsample of 15 SNe assumed to have suffered little or no extinction in their host galaxies was used to measure color excesses for the whole sample. By plotting the ratios of color excesses involving optical and NIR filters and comparing the values predicted by the CCM+O extinction law for interstellar dust, a value of the total-to-selective absorption coefficient of $R_V \\approx 1.7$ was derived, which is significantly lower than the Galactic average of $3.1$. However, this value is largely influenced by two very red objects in the sample, SNe 2005A and 2006X. When the same calculations are repeated after excluding these two objects, a value of $R_V=3.2\\pm0.4$ is found, in agreement with the Galactic average. An alternative way of estimating $R_V$ is to express the absolute magnitudes of the SNe as a two-parameter function of the decline rate and color, and then to use the Hubble-flow distances (or Cepheid and SBF distances) of the SNe to derive the best fit parameters through $\\chi^2$ minimization. This method was first employed by \\citet{tripp98} and \\citet{tripp99}, who related the absolute magnitude in $B$ to the $B-V$ color and derived values of $R_V \\approx$ 1--1.5. When this same technique is applied to the CSP sample of SNe~Ia using the peak magnitudes in $B$ and $J$ bands and the $B-V$ and $V-J$ colors, a value of $R_V \\sim$1--2 is obtained, regardless of whether the heavily reddened events, SNe 2005A and 2006X, are included or excluded. An equivalent analysis can also be carried out relating absolute magnitude to the decline rate and color excess. Performing such fits for all available bands ($ugriBVYJHK_s$) again leads to systematically low values of the CCM+O-law parameter ($R_V \\approx 1.5$) regardless of whether the two highly reddened SNe are included in the sample. These conflicting results on the value of $R_V$ reflect the variety of results found in the literature on this subject. Most early attempts to derive the average properties of the host-galaxy reddening indicated that $R_V$ was unusually low \\citep[see][and references therein]{branch92}. However, the available light curves at the time were largely photographic and there was considerable additional uncertainty regarding the nature of the intrinsic colors of SNe~Ia at maximum light. The first modern study based on CCD data was made by \\citet{riess96}, who used the ratios of color excesses measured in $B-V$, $V-R$, and $V-I$ for a sample of 20 SNe~Ia to derive a value of $R_V = 2.55\\pm0.30$. This analysis differed from previous studies in that the intrinsic color variation as a function of decline rate was taken fully into account in deriving the color excesses. Similar values of $R_V$ have been found by \\citet{phillips99}, \\citet{altavilla04}, \\citet{reindl05}, and \\citet{wangx06}. However, beginning with \\citet{tripp98}, studies based on the technique of minimizing the dispersion in the Hubble diagram of SNe~Ia have nearly uniformly led to values of $R_V$ in the range of $\\sim$1--2 \\citep[e.g.,][]{tripp99, astier06, conley07}. If there is a pattern to these results, including our own in this paper, it would seem that attempts to measure $R_V$ via comparison of colors or color excesses tend to give larger values than the procedure of minimizing the scatter in the Hubble diagram with $R_V$ treated as a free parameter. Nevertheless, there are exceptions to this rule such as the recent paper by \\citet{nobili08}, who studied the color evolution in $U-B$, $B-V$, $V-R$, and $R-I$ of 69 SNe~Ia with moderate reddening and obtained a value of $R_V = 1.01\\pm0.25$. Common sense suggests that at least some portion of the observed reddening of SNe~Ia in spiral galaxies like the Milky Way must be due to dust in the interstellar medium of these galaxies. Although only a few studies have been carried out of the nature of the reddening law in external galaxies, these are generally consistent with $R_V \\approx 3$, albeit with a large dispersion \\citep{goudfrooij94,patil07,ostman08}. As shown in \\S~\\ref{sec:cered}, the ratios of the measured color excesses involving optical and NIR bands with respect to $E(B-V)$ are consistent with values of $R_V \\approx 3$. As discussed by \\citet{krisciunas07}, inclusion of the NIR bands in this type of analysis improves considerably the accuracy with which $R_V$ can be measured. Hence, it is tempting to conclude that we are measuring reddening produced by dust with very similar properties to that found in the interstellar medium of the disk of the Milky Way. It will be interesting to see if this finding holds up as many more SNe~Ia with reliable optical$-$NIR color excesses are added to the CSP sample. If we are, indeed, observing host-galaxy reddening due to ``normal'' Galactic-type dust, then the fact that low values of $R_V$ ($\\sim$1--2) are obtained when it is treated as a free parameter in minimizing the dispersion in the Hubble diagram is puzzling. The implication is that there is an intrinsic dispersion in the colors of SNe~Ia which is correlated with luminosity, but is independent of the decline rate. Our finding in \\S~\\ref{sec:cered} that there is a small but systematic difference between color-excess measurements, $E(B-V)$, made at late epochs using the Lira law and those derived from the maximum light magnitudes, and that this difference correlates with color and perhaps also with absolute magnitude, may well be related to this. Very red SNe potentially allow the dust reddening law to be studied on an individual basis. Including the results for SN~2005A given in this paper, there are now six SNe~Ia for which a determination of $R_V$ has been made from optical and NIR photometry. The results for these objects are summarized in Table~\\ref{tab:redsne}. A weighted mean of the six determinations gives $R_V = 1.6$, with a surprisingly small rms of 0.3. The relatively narrow range of decline rates is also noteworthy, averaging $\\dm = 1.2$ with a dispersion of only 0.1. Except for SN~2001el, the color excesses for these SNe are all very large. \\citet{jha07} found that the distribution of $E(B-V)$ for SNe~Ia is well approximated by an exponential function with $\\tau = 0.138 \\pm 0.023$~mag. Hence, values of $E(B-V) > 1$ mag should be extremely rare, although it must be kept in mind that the sample of SNe analyzed by Jha et al. was culled from several sources with different selection criteria. Assuming that the Jha et al. distribution is correct, less than one SN for every thousand observed would be expected to have such a large amount of reddening --- yet according to the Asiago Supernova Catalog\\footnote[15]{http://web.oapd.inaf.it/supern/cat/ .}, the total number of nearby ($z < 0.02$) SNe~Ia discovered from 1985 through 2008 was only 358. The fact that these heavily reddened SNe are much more common than expected, combined with the remarkable similarity of decline rates and $R_V$ values for the five objects in Table~\\ref{tab:redsne} with $E(B-V) > 1$ mag, leads us to speculate that these events actually represent a physically distinct subclass of SNe~Ia. Hence, while extremely interesting in themselves, these objects like SNe 2005A and 2006X are not representative of the class of SNe~Ia employed in cosmological studies. \\cite{wang05} has suggested that the presence of circumstellar dust distributed in a shell around the SN can produce reddening compatible with the observed low values of $R_V$. This hypothesis is supported by our observations of SNe~2005A and 2006X. Specifically, we find that the red colors of these two SNe are better matched by a model where multiple scattering of photons by circumstellar dust steepens the effective extinction law \\citep{goobar08}, than by the standard CCM+O Galactic reddening law with a low value of $R_V$. Our results on the absolute peak magnitude calibrations for this new sample of CSP SNe confirm those obtained in previous studies, while extending the calibration to the optical $ugri$ and the NIR $YJHK_s$ bands. Fits to the subsample of best-observed SNe employing the \\citet{tripp98} model, which assumes the absolute magnitudes to be a two-parameter function of the decline rate and color, give rms dispersions of 0.12--0.15~mag. Remarkably, the same fits are found to apply equally well to either highly-reddened objects such as SNe 2005A and 2006X, or fast-declining, intrinsically-red events like 2005ke and 2006mr. The simplicity of this method is appealing. We have also carried out fits which assume the absolute peak magnitudes to be a function of the decline rate and the measured color excesses. These give quite similar results to those obtained with the Tripp model. Fits to a large number of filters and colors give dispersions in the range of 0.12--0.20~mag for SNe with $0.7 < \\dm <1.7$ mag. The quality of our NIR light curves allows, for the first time, detection of a weak dependence of the $J$-band luminosity on decline rate. Combining the calibrations from all bands, we created a single Hubble diagram for the 23 best-observed SNe. Although the resulting scatter of $0.12$~mag (6\\% in distance) is excellent, it does not represent a significant improvement over the precision obtained with individual filters. This is attributed to the fact that, for any particular SN, there is a significant correlation in the error in the distance moduli obtained using different filters. The source of these correlated errors appears to be the peculiar velocities of the SN host galaxies. If true, this implies that the derived distances to the SNe are actually precise to 3--4\\%, making SNe~Ia competitive with even Cepheid variables as extragalactic distance indicators. Finally, the set of NIR light curves presented here confirms the advantages of working in this wavelength region for cosmological studies. Fit~8 of Table~\\ref{tab:tb} and Fit~13 of Table~\\ref{tab:ld} show that using either the $V-J$ color or color excess to correct the $J$ absolute magnitudes of the best-observed SNe (excluding the fast decliners) yields a dispersion of only 0.12~mag. Moreover, this result is insensitive to the exact form of the reddening law since the extinction corrections in the NIR are small. Note that the values derived for $\\sigma_{rm SN}$ of 0.02--0.04~mag suggest that the intrinsic dispersion in the $J$ band may be quite small. This is obviously a preliminary result which must be treated with caution --- however, we will soon have a sample of $\\sim$80 SNe~Ia to test this. If confirmed, it implies that if the observational uncertainties can be reduced, there is much to be gained by extending the rest-wavelength coverage of future dark-energy experiments to include the $J$ band." }, "0910/0910.3910_arXiv.txt": { "abstract": "We use our model for the formation and evolution of galaxies within a two-phase galaxy formation scenario, showing that the high-redshift domain typically supports the growth of spheroidal systems, whereas at low redshifts the predominant baryonic growth mechanism is quiescent and may therefore support the growth of a disc structure. Under this framework we investigate the evolving galaxy population by comparing key observations at both low and high-redshifts, finding generally good agreement. By analysing the evolutionary properties of this model, we are able to recreate several features of the evolving galaxy population with redshift, naturally reproducing number counts of massive star-forming galaxies at high redshifts, along with the galaxy scaling relations, star formation rate density and evolution of the stellar mass function. Building upon these encouraging agreements, we make model predictions that can be tested by future observations. In particular, we present the expected evolution to z=2 of the super-massive black hole mass function, and we show that the gas fraction in galaxies should decrease with increasing redshift in a mass, with more and more evolution going to higher and higher masses. Also, the characteristic transition mass from disc to bulge dominated system should decrease with increasing redshift. ", "introduction": " ", "conclusions": "" }, "0910/0910.1915_arXiv.txt": { "abstract": "The single-degenerate channel is widely accepted as the progenitors of type Ia supernovae (SNe Ia). Following the work of Meng, Chen \\& Han (2009), we reproduced the birth rate and age of supernovae like SN 2006X by the single-degenerate model (WD + MS) with an optically thick wind, which may imply that the progenitor of SN 2006X is a WD + MS system. ", "introduction": "\\label{sect:1} Although type Ia supernovae (SNe Ia) showed their importance in determining cosmological parameters, e.g. $\\Omega_{\\rm M}$ and $\\Omega_{\\Lambda}$ ((Riess et al. 1998; Perlmutter et al. 1999), the progenitor systems of SNe Ia have not been confidently identified yet (Hillebrandt \\& Niemeyer 2000; Leibundgut 2000). Two basic scenarios for the progenitor of SN Ia have been discussed over the last three decades. One is a single degenerate (SD) model (Whelan \\& Iben 1973; Nomoto, Thielemann \\& Yokoi 1984), in which a CO WD increases its mass by accreting hydrogen- or helium-rich matter from its companion, and explodes when its mass approaches the Chandrasekhar mass limit. The companion may be a main-sequence star (WD + MS) or a red-giant star (WD + RG) (Yungelson et al. 1995; Li \\& van den Heuvel 1997; Hachisu et al. 1999a,b; Nomoto et al. 1999, 2003; Langer et al. 2000; Han \\& Podsiadlowski 2004; Chen \\& Li 2007, 2009; Han 2008; Meng, Chen \\& Han 2009; L\\\"{u} et al. 2009, Wang et al. 2009a, b). An alternative is the double degenerate (DD) model (Iben \\& Tutukov 1984; Webbink 1984), in which a system consisting of two CO WDs loses orbital angular momentum by gravitational wave radiation and merges finally. The merger may explode if the total mass of the system exceeds the Chandrasekhar mass limit (see the reviews by Hillebrandt \\& Niemeyer 2000 and Leibundgut 2000). In theory, a large amount of circumstellar materials (CSM) may form around SNe Ia via an optically thick wind for the SD model (Hachisu et al. 1996), while there is no CSM around DD systems. Then, a basic method to distinguish the two progenitor models is to find the CSM around progenitor systems. The CSM may play a key role to solve the problem of the low value of reddening ratio (Wang 2005), which is very important for precision cosmology (Wang et al. 2008). In addition, it is possible for the CSM to be the origin of color excess of SNe Ia (Meng et al. 2009). Evidence for CSM was first found in SN2002ic (Hamuy et al. 2003), which has shown extremely pronounced hydrogen emission lines which have been interpreted as a sign of strong interaction between supernova ejecta and CSM. The discovery of SN2002ic may uphold the SD model (Han \\& Podsiadlowski 2006). The evidence for CSM was also found in a normal SN Ia (SN 2006X) defined by Branch, Fisher \\& Nugent (1993) and the CSM is proposed to be from a wind from a red-giant companion (Patat et al. 2007), while Hachisu et al. (2008) argued a WD + MS nature for this SN Ia. Blondin et al. (2009) found a similar signal to that of SN 2006X in another SNe Ia (SN 1999cl) and their results indicated that the supernovae like SN 2006X are rare objects ($2/31\\sim6\\%$). Recently, the third example like SN 2006X (SN 2007le) was reported (Simon et al. 2009), and then the ratio of 2006X-like supernova is increased to $3/32\\sim9\\%$. In theory, SNe Ia can explode during the optically thick wind phase or after the optically thick wind while in stable or unstable hydrogen-burning phase (Han \\& Podsiadlowski 2004). The materials lost as the optically thick wind form CSM (Hachisu et al. 2008; Meng et al. 2009). If a SN Ia explodes after the optically thick wind, the CSM have been dispersed too thin to be detected immediately after the SN Ia explosion (Hachisu et al. 2008; Meng, Yang \\& Geng 2009). If a SNe Ia explodes during the optically thick wind phase, the materials lost from system form CSM very near the SN Ia, which may show the signal similar to SN 2006X (Hachisu et al. 2008; Meng, Yang \\& Geng 2009). Recently, Meng, Chen \\& Han (2009) performed binary stellar evolution calculations for more than 25,000 close WD + MS binary systems with various metallicities. In their works, the prescription of Hachisu et al. (1999a) for the accretion efficiency of hydrogen-rich material was adopted by assuming an optically thick wind (Hachisu et al. 1996), and then their works provide a possibility to check whether the SD model with optically thick wind can explain the birth rate of supernovae like SN 2006X. The purpose of this paper is to check the possibility by a binary population synthesis (BPS) approach. In section \\ref{sect:2}, we simply describe our method. We show the results in section \\ref{sect:3} and give discussions and conclusions in section \\ref{sect:4}. ", "conclusions": "\\label{sect:4} \\subsection{the age of SN 2006X} \\label{sect:4.1} SN 2006X is the first case that show a variable Na I D line in its spectral, which clearly shows a signal of CSM (Patat et al. 2007). Patat et al. (2007) also noticed a relatively low expansion velocity of the CSM ($\\sim50$ ${\\rm km/s}$). The low expansion velocity seems to imply that the progenitor of SN 2006X belongs to a WD + RG systems. However, Hachisu et al. (2008) showed that under the assumption of the optically thick wind, the WD + MS system also may produce the low-velocity CSM. In this paper, we show that if SN 2006X is from a WD + MS system with the optically thick wind, the age of its progenitor is smaller than 1 Gyr, which implies that there is star formation during the recent 1 Gyr in its host galaxy. The host galaxy of SN 2006X, NGC 4321 (M100), is a well-studied spiral galaxy (e.g. Ho et al. 1997) and SN 2006X locates near one of its arms (Wang et al. 2008). As one of the largest spiral galaxies in the Virgo cluster, it produced SNe 1901B, 1914A, 1959E, 1979C, and 2006X in roughly the last century and shows significant signal of star formation at present (Kanpen et al. 1993, 1996). Recently, Blondin et al. (2009) found a similar signal to that of SN 2006X in another SNe Ia (SN 1999cl) in archives. Its host galaxy, NGC4501 (M88), is also a spiral galaxy and SN 1999cl locates at one of its arms (Krisciunas et al. 2000). During the last 1 Gyr, the star formation is significant in the galaxy (Wong \\& Blitz 2002). The host galaxy of SN 2007le (NGC 7721) is a Sc galaxy, and a recent star formation is also expected (Iglesias-P\\'{a}ramo et al. 2006; Simon et al. 2009). So, all the supernovae fulfill the age constraint from WD + MS system. \\subsection{Progenitor system} \\label{sect:4.2} Following the study of Meng, Chen \\& Han (2009), in this paper, we show the evolution of birth rate of supernovae like SN 2006X by assuming that if a SN Ia explodes at the optically thick wind phase, it is a 2006X-like supernova. Based on the assumption, we can reproduce the birth rate of 2006X-like objects. Please keep in mind that we only consider the case of WD + MS channel in this paper. However, the progenitor of SN 2006X is still an open question. Patat et al. (2007) suggested that the progenitor of SN 2006X belongs to WD + RG system such a RS Oph based on the velocity of absorptions line of Na I D. L\\\"{u} et al. (2009) designed a WD + RG model with an aspherical stellar wind and equatorial disk, and then they may explain some properties of SN 2006X. However, the birth rate of 2006X-like objects obtained in L\\\"{u} et al. (2009) is too low (less than 1\\%) to compare with that observed. Based on the assumption of the optically thick wind and a mass-stripping effect, Hachisu et al. (2008) argued that the progenitor of SN 2006X should be a WD + MS system and they also can well explain the properties of SN 2006X. Although the treatment of binary evolution in Meng, Chen \\& Han (2009) is different from that in Hachisu et al. (2008), Meng, Chen \\& Han (2009) obtained similar results for the SNe Ia exploding at the optically thick wind phase, e.g. there is no supernova explosion at the optically thick wind phase when the initial mass of CO WD is smaller than 0.9 $M_{\\odot}$. So, considering the results in this paper, all the properties of SN 2006X can be explained by the WD + MS channel with the optically thick wind, including its birth rate and its age. Then, we suggest that the progenitor of supernovae like SN 2006X is WD + MS system, not WD + RG system." }, "0910/0910.4192_arXiv.txt": { "abstract": "The Fermi Gamma-ray Space Telescope (FGST) has opened a new high-energy window in the study of Gamma-Ray Bursts (GRBs). Here we present a thorough analysis of GRB~080825C, which triggered the Fermi Gamma-ray Burst Monitor (GBM), and was the first firm detection of a GRB by the Fermi Large Area Telescope (LAT). We discuss the LAT event selections, background estimation, significance calculations, and localization for Fermi GRBs in general and GRB~080825C in particular. We show the results of temporal and time-resolved spectral analysis of the GBM and LAT data. We also present some theoretical interpretation of GRB~080825C observations as well as some common features observed in other LAT GRBs. ", "introduction": "Gamma-Ray Bursts (GRBs) originate from the most luminous explosions in the universe and more than 35 years after their discovery in 1967 \\citep{kel73}, many questions remain to be answered about their possible progenitors, the composition of the ultra-relativistic outflows that power them, and the dominant emission mechanism for their prompt gamma rays. The Burst And Transient Source Experiment (BATSE) onboard the Compton Gamma-Ray Observatory (CGRO; 1991-2000) made significant advances to the field, thoroughly exploring the 25~keV -- 2~MeV energy range with detailed population studies of the prompt gamma-ray emission. Burst spectra were found to be well described by the Band function \\citep{band93}, which consists of two smoothly connected power laws. It was understood, however, that observations of GRBs at higher energies were of crucial importance to resolve some of the open issues: constrain the bulk Lorentz factor of the outflow and the distance from the central source to the gamma-ray emission region, distinguish between hadronic and leptonic origins of the gamma-ray emission, and probe for signatures of Ultra High Energy Cosmic Rays (UHECRs) which could be accelerated within GRB jets \\citep[see][for a review of the prospects for GRB science with Fermi LAT]{band08}. Constraints on the origin of the high-energy emission from GRBs are quite limited due to both the small number of bursts with firm high-energy detection and the small number of events that were detected in such cases. High-energy emission from GRBs was first observed by the Energetic Gamma-Ray Experiment Telescope (EGRET, covering the energy range from $30\\;$MeV to $30\\;$GeV) onboard CGRO. Emission above $100\\;$MeV was detected in five cases: GRBs~910503, 910601, 930131, 940217 and 940301 \\citep{dingus95}. One of these sources, GRB~930131, had high-energy emission that was consistent with an extrapolation from its spectrum obtained with BATSE between 25~keV -- 4~MeV~\\citep{som94}. In contrast, evidence for an additional high-energy component up to $200\\;$MeV with a different temporal behavior to the low-energy component was discovered in GRB~941017 \\citep[in EGRET's calorimeter TASC; ][]{gon03}. The high-energy emission for the latter GRB lasted more than 200 seconds with a single spectral component being ruled out. A unique aspect of the high-energy emission in GRB~940217 was its duration, which lasted up to $\\sim$$90$ minutes after the BATSE GRB trigger, including an 18 GeV photon at $\\sim$$75$ minutes post-trigger \\citep{hur94}. More recently, the GRID instrument onboard Astro-rivelatore Gamma a Immagini LEggero (AGILE) detected 10 high-energy events with energies up to $300\\;$MeV from GRB 080514B, in coincidence with its lower energy emission, with a significance of $3.0\\;\\sigma$ \\citep{giu08}. The Fermi Gamma-ray Space Telescope was launched on June 11 2008 and provides an unprecedented energy coverage and sensitivity for the study of high-energy emission in GRBs. It is composed of two instruments: the Gamma-ray Burst Monitor \\citep[GBM;][]{meg08} and the Large Area Telescope \\citep[LAT;][]{atw09}. The GBM covers the entire unocculted sky with 12 sodium iodide (NaI) detectors with different orientation placed around the spacecraft and covering an energy range from 8 keV to 1 MeV, and two bismuth germanate (BGO) scintillators placed on opposite sides of the spacecraft with energy coverage from 200 keV to 40 MeV. The LAT is a pair conversion telescope made up of $4 \\times 4$ arrays of silicon strip trackers and cesium iodide calorimeter modules covered by a segmented anti-coincidence detector designed to efficiently reject charged particles. The energy coverage of the LAT instrument ranges from 20 MeV to more than 300~GeV with a field-of-view (FoV) of $\\sim$$2.4$ steradians. Note that the LAT effective area is still non-zero even as far out as 70 degrees off-axis which allows the detection of bursts with such high incident angles. As of June 1 2009, 9 GRBs have been detected by the LAT at energies above 100~MeV: GRB~080825C \\citep{bou08}, GRB~080916C \\citep{abd09, tajima08}, GRB~081024B \\citep{omo08}, GRB~081215A \\citep{mcenery08}, GRB~090217 \\citep{ohn09}, GRB~090323 \\citep{ohnovdh09}, GRB~090328 \\citep{mcenery09, cutini09}, GRB~090510 \\citep{ohnopela09, omo09}. In this paper, we report the observations and analyses of gamma-ray emission from GRB~080825C, the first GRB detected by both the GBM and the LAT instruments. Section \\ref{sec:observations} will present the GBM and LAT observations along with the various methods used for data analysis, section \\ref{sec:dataana} provides the results of detailed time-resolved spectroscopy, and section \\ref{sec:discussion} discusses the theoretical interpretation of our observations and compares the properties of this event to the ones observed in some other LAT GRBs. ", "conclusions": "GRB~080825C is the first GRB detected by the LAT, with 13 events above 80~MeV and a detection significance of $\\sim$$6\\;\\sigma$. The highest energy events, up to $\\sim$$600\\;$MeV, are detected at late times, $\\sim$$25-35\\;$s after the GBM trigger, when the emission in GBM has decreased close to the background level. The lack of $>1$~GeV events in the LAT and the large angle of the source to the LAT boresight result in a localization uncertainty of $1.1^\\circ$ (statistical plus systematic) at the $1\\;\\sigma$ level. The prompt emission spectrum from both instruments onboard Fermi covers over 5 decades in energy. We have performed time-resolved spectral analysis using the two GBM NaI detectors with the brightest GRB signal, both GBM BGO detectors, and for the LAT with an event selection scheme that is optimized for GRB analysis. We have carefully taken the energy-dependent backgrounds into account for both GBM and LAT, and studied the systematic uncertainties in the spectral analysis. The time-integrated and time-resolved spectra are well fit by the Band function with a hard-to-soft evolution in the first $25\\;$s: $E_{\\rm{peak}}$ evolves from $\\sim$$300$ to $\\sim$$150\\;$keV, the high-energy power-law index $\\beta$ is constant at a value of $\\sim$$-2.5$, while the low-energy power-law index $\\alpha$ is fairly constant except for the second time bin which contains the first LAT events. In the last time bin, $\\sim$$25-35\\;$s after the GBM trigger time, the GBM data are barely above background level, and the spectrum is best fit by a single power law with an index of $\\sim$$-2$ which is significantly harder than the $\\beta$ values of the earlier intervals. The duration and start time of the late broad peak in the high-energy emission, $\\sim$$16-31\\;$s after the trigger, suggest that this peak is emitted by the external reverse or forward shocks, rather than by internal dissipation within the GRB outflow (e.g., internal shocks or magnetic reconnection). The relatively fast flux decay after this peak slightly favors a reverse-forward shock `external' Compton origin over a forward shock SSC origin. Although the origin and emission mechanism for this late peak cannot be conclusively determined because of low number statistics (and the lack of observations at X-ray or optical wavelengths, due to the poor GRB localization), the external shock origin is further supported by the change in spectral behavior, in particular of the spectral index, at these late times. Observations of more, brighter GRBs with both GBM and LAT will be able to test this hypothesis." }, "0910/0910.4471_arXiv.txt": { "abstract": "{% Nova V5116 Sgr 2005 No. 2, discovered on 2005 July 4, was observed with XMM-Newton in March 2007, 20 months after the optical outburst. The X-ray spectrum showed that the nova had evolved to a pure supersoft X-ray source, indicative of residual H-burning on top of the white dwarf. The X-ray light-curve shows abrupt decreases and increases of the flux by a factor 8 with a periodicity of 2.97~h, consistent with the possible orbital period of the system. The EPIC spectra are well fit with an ONe white dwarf atmosphere model, with the same temperature both in the low and the high flux periods. This rules out an intrinsic variation of the X-ray source as the origin of the flux changes, and points to a possible partial eclipse as the origin of the variable light curve. The RGS high resolution spectra support this scenario showing a number of emission features in the low flux state, which either disappear or change into absorption feature in the high flux state. A new XMM-Newton observation in March 2009 shows the SSS had turned off and V5116 Sgr had evolved into a weaker and harder X-ray source.} ", "introduction": "V5116 Sgr was discovered as Nova Sgr 2005b on 2005 July 4.049~UT, with magnitude $\\sim$8.0, rising to mag 7.2 on July 5.085 \\cite{lil05}. It was a fast nova ($t_3=39$ days) belonging to the Fe II class in the Williams (1992) classification (Williams et al. 2008). The expansion velocity derived from a sharp P-Cyg profile detected in a spectrum taken on July 5.099 was $\\sim$1300~km/s. IR spectroscopy on July 15 showed emission lines with FWHM~$\\sim$2200~km/s \\cite{rus05}. Photometric observations obtained during 13 nights in the period August-October 2006 show the optical light curve modulated with a period of $2.9712\\pm0.0024$~h \\cite{dob07}, which the authors interpret as the orbital period. They propose that the light-curve indicates a high inclination system with an irradiation effect on the secondary star. The estimated distance to V5116~Sgr from the optical light curve is $11\\pm3$~kpc \\cite{sal08}. A first X-ray observation with Swift/XRT (0.3--10~keV) in August 2005 yielded a marginal detection with 1.2($\\pm$1.0) $\\times10^{-3}$~cts/s \\cite{nes07a}. Two years later, on 2007 August 7, Swift/XRT showed the nova as a SSS, with 0.56 $\\pm$0.1~cts/s \\cite{nes07b}. A first fit with a blackbody indicated $T\\sim4.5\\times10^5K$. A 35~ks Chandra spectrum obtained on 2007 August 28 was fit with a WD atmospheric model with N$_{\\rm H}=4.3\\times10^{21}$~cm$^{-2}$ and $T=4.65\\times10^5K$ \\cite{NelOri07}. ", "conclusions": "" }, "0910/0910.3856_arXiv.txt": { "abstract": "We discuss three effects of axial rotation at low metallicity. The first one is the mixing of the chemical species which is predicted to be more efficient in low metallicity environments. A consequence is the production of important quantities of primary $^{14}$N, $^{13}$C, $^{22}$Ne and a strong impact on the nucleosynthesis of the {\\it s}-process elements. The second effect is a consequence of the first. Strong mixing makes possible the apparition at the surface of important quantities of CNO elements. This increases the opacity of the outer layers and may trigger important mass loss by line driven winds. The third effect is the fact that, during the main-sequence phase, stars, at very low metallicity, reach more easily than their more metal rich counterparts, the critical velocity\\footnote{The critical velocity is the surface equatorial velocity such that the centrifugal acceleration compensates for the local gravity.}. We discuss the respective importance of these three effects as a function of the metallicity. We show the consequences for the early chemical evolution of the galactic halo and for explaining the CEMP stars. We conclude that rotation is probably a key feature which contributes in an important way to shape the evolution of the first stellar generations in the Universe. ", "introduction": "The study of the most iron poor stars in the halo offers a unique window on the nature and the evolution of the first generations of stars which provided the matter from which these halo stars were formed. Constraints on the first stellar generations may provide interesting views on topical questions as the nature of the stars which reionize the Universe, the initial mass function of the first stellar generations, the frequency of long soft Gamma Ray Bursts at very low metallicity, the timescale for the mixing of the newly synthesized products with interstellar medium material, or the conditions of star formation in the very early life of our Galaxy. \\begin{figure}[t] \\includegraphics[width=2.4in,height=2.4in,angle=0]{chem5-2.eps} \\hfill \\includegraphics[width=2.4in,height=2.4in,angle=0]{chem5N-1.eps} \\caption{{\\it Left panel} : variation of the abundances of various elements (in mass fraction) as a function of the Lagrangian mass in the outer layers of a 60 M$_\\odot$ model at $Z=10^{-5}$ with $\\upsilon_{\\rm ini}=800$ km s$^{-1}$ at the end of the core He-burning phase. The grey area covers the CO core. {\\it Right panel :} same as left part for a non rotating 60 M$_\\odot$ model at $Z=10^{-5}$. Models computed by \\citet{paperVIII2002} and \\citet{Meynet2006}.} \\label{struc} \\end{figure} In this context, many observed features of the very iron poor stars are quite surprising and were not at all expected: \\begin{itemize} \\item One expected that halo stars, having a very low [Fe/H], should show an important scatter in their abundances. Indeed, at these very early times, stars are expected to form from not well mixed clumps and thus should bear the nucleosynthetic signatures of a few peculiar events. But \\citet{Cayrel2004} find that the scatter in the abundances of several elements ratios (e.g. [Cr/Fe]) is very small. It can be as low as 0.05 dex. In contrast, r-process elements and nitrogen for instance reveal large star-to-star scatters \\citep{Ryan1996, Honda2004, Andrievsky2009}. Explanation are proposed in \\citet{Ishimaru2004, Ishimaru2006, Cescutti2008}. \\item One expected that the very first generations of Pop. III stars were very massive and contributed to the enrichment of the interstellar medium through Pair Instability Supernovae \\citep[PISN,][]{Barkat1967, Bond1984}. The nucleosynthetic pattern produced by such events is well understood and presents specific features \\citep{HegerWoosley2005}, which were expected to be observed at least in some iron-poor halo stars. At the moment no trace of PISN has been found \\citep{Cayrel2004}. Some authors have suggested an explanation why this signature has, up to now, escaped detection \\citep{Karlsson2008, Ekstrom2008d}. \\item Spectroscopic observations of halo stars \\citep[e.g.][]{Spite2005} indicate a primary production of nitrogen over a large metallicity range at low metallicity. According to the chemical evolution models of \\citet{Chiappini2006, Chiappini2008}, primary nitrogen production by non-rotating or slow-rotating Pop III stars is not sufficient to explain the observations. For instance, primary nitrogen needs to be produced on larger ranges of masses and metallicities than expected in non rotating standard models. A similar plateau is observed in Damped Lyman Alpha (DLA) systems \\citep{Pettini2008}. \\item Halo stars with log(O/H)+12 inferior to about 6.5 present higher C/O ratios than halo stars with log(O/H)+12 between 6.5 and 8.2 \\citep{Akerman2004, Spite2005}. Again a similar trend is observed in DLAs \\citep{Pettini2008}. This is not predicted by slow rotating models. \\item The features listed above concern the bulk of the halo field stars. Now in addition to this ``normal'' population, there exists a group of stars showing a very different composition. These stars are collectively called Carbon-Enhanced Metal Poor stars (CEMP) and present high [C/Fe] ratios \\citep[see the review by][]{Beers2005}. The scatter in [C/Fe] is quite important indicating that these stars were not formed from a well mixed reservoir but acquired their peculiar high carbon abundance from locally enriched material. Some of these stars show also strong overabundances (with respect to iron) of nitrogen, oxygen and other heavy elements. At first order, these stars do appear to be formed from material having been processed mainly by H- and He-burning processes. \\item Very interestingly, halo stars in clusters also present some surprising features. While part of the cluster population follows the same trend as the ``normal'' population of the field, another part presents very different chemical patterns, with for instance, strong depletions of oxygen and strong enhancements in sodium \\citep[see the review by][]{Gratton2004}. These stars do appear to be formed from material having been processed only by H burning. \\end{itemize} Valid models should provide explanations for all these features. Ideally they should also provide some predictions for not yet observed characteristics. In the following we discuss the possible role of stellar axial rotation and we show how it can deeply affect the nucleosynthetic outputs of stars both in the massive ($>$ 8 M$_\\odot$) and intermediate mass range (2 $<$ M/M$_\\odot$ $<$ 8) and how it can explain some of the above observed features. ", "conclusions": "" }, "0910/0910.0855_arXiv.txt": { "abstract": "In this paper we report on observations of the CoRoT LRa1 field with the Berlin Exoplanet Search Telescope (BEST). The current paper is part of the series of papers describing the results of our stellar variability survey. BEST is a small aperture telescope with a wide field-of-view (FOV). It is dedicated to search for stellar variability within the target fields of the CoRoT space mission to aid in minimizing false-alarm rates and identify potential targets for additional science. The LRa1 field is CoRoT's third observed field and the second long run field located in the galactic anticenter direction. We observed the LRa1 stellar field on $23$ nights between November and March $2005/2006$. From $6099$ stars marked as variable, $39$ were classified as periodic variable stars and $27$ of them are within the CoRoT FOV. We also confirmed the variability for $4$ stars listed in GCVS catalog. ", "introduction": "In previous articles belonging to a series dedicated to the variability survey in the CoRoT observational fields with BEST we presented the results of our survey on variable stars in the CoRoT IRa1~\\cite{Kabath07} and LRc1~\\cite{Karoff07} observational fields. The fields were observed with BEST~\\cite{Rauer04}, a small aperture telescope with a wide field-of-view (FOV) developed and operated by the Institut f\\\"{u}r Planetenforschung of Deutsches Zentrum f\\\"ur Luft- und Raumfahrt (DLR). As the BEST magnitude range covers approximately that of the CoRoT it is thus well suited for ground based support of the CoRoT space mission~\\cite{Baglin98}. CoRoT was launched in December $2006$. The scientific goals of the mission are observations of selected stellar fields in order to find transiting extrasolar planets and to perform astroseismology of selected stars. The duration of the mission is planned to be $2.5$ years with $4$ long ($150$ days), $4$ short (up to $30$ days) and one initial ($30$ days) run. The CCD camera of CoRoT is divided into four segments from which two are dedicated to a transiting extrasolar planets survey and the other two to the asteroseismic survey. The total FOV covers $2.7^\\circ \\times 3.05^\\circ$. The point spread function (PSF) of CoRoT's extrasolar planet survey is about $80$ pixels for a $13$ mag K2 star and the PSF of the astroseismology field is about $410$ pixels for a solar type star at magnitude $5.7$ see~\\cite{Boisnard2006}. The prisms mounted in front of CoRoT's CCD provide colour information on observed objects which can help to determine the type of variability of the central star. The BEST survey telescope system is used to discover and characterize variable stars in the CoRoT fields prior to CoRoT observations. The observational data can also be used as a complimentary information to the incoming CoRoT data. In addition BEST observations should also point out potentially interesting objects for the CoRoT additional science programs, such as binaries, $\\delta$ Scuti stars etc.. The advantage of our observations is the long time line which can provide extended lightcurves for several thousands of stars in the comparable magnitude range as CoRoT data and thus better understanding of e.g. potential variation of the lightcurve. The resulting information about the variability in the stellar fields observed by BEST will be provided to the scientific community via the BEST archive linked to CoRoT's EXODAT~\\cite{Deleuil2006} database. ", "conclusions": "We performed photometry on CoRoT LRa1 field with BEST telescope to detect stellar variability. We identified $6099$ stars which were marked as variable and $44$ of them are showing a regular period. In total we detected $39$ new variable stars. The period ranges are usually between $0.1$ $<$ P $<$ $3$ days, however two new longperiodic stars with periods about $7$ and $21$ days were detected. The relatively small period range is given due to limited data set and due to duty cylce coverage which was disturbed by the bad weather period in December. We also confirm the period for the stars GU Mon, DD Mon and V 404 Mon which are already listed in the GCVS catalogue. The period for the GCVS star V 501 Mon could not be confirmed because no maxima/minima are present in our data set for that star. The star CD Mon is a longperiodic Mira type star and is also present in our data set. The newly found variable stars are currently observed within CoRoT's additional science programs. The information provided with our survey brings an additional information when analysing the CoRoT data and may avoid a false alarm for CoRoT transiting planet candidates. We gladly provide additional information about our data archive upon a request." }, "0910/0910.5641_arXiv.txt": { "abstract": "{Realistic stellar models are essential to the forward modelling approach in asteroseismology. For practicality however, certain model assumptions are also required. For example, in the case of subdwarf B stars, one usually starts with zero-age horizontal branch structures without following the progenitor evolution. } {We analyse the effects of common assumptions in subdwarf B models on the $g$-mode pulsational properties. We investigate if and how the pulsation periods are affected by the H-profile in the core-envelope transition zone. Furthermore, the effects of C-production and convective mixing during the core helium flash are evaluated. Finally, we reanalyse the effects of stellar opacities on the mode excitation in subdwarf B stars.} {We computed detailed stellar evolutionary models of subdwarf B stars, and their non-adiabatic pulsational properties. Atomic diffusion of H and He is included consistently during the evolution calculations. The number fractions of Fe and Ni are gradually increased by up to a factor of 10 around $\\log T=5.3$. This is necessary for mode excitation and to approximate the resulting effects of radiative levitation. We performed a pulsational stability analysis on a grid of subdwarf B models constructed with OPAL and OP opacities.} {We find that helium settling causes a shift in the theoretical blue edge of the $g$-mode instability domain to higher effective temperatures. This results in a closer match to the observed instability strip of long-period sdB pulsators, particularly for $ l\\leq3$ modes. We show further that the $g$-mode spectrum is extremely sensitive to the H-profile in the core-envelope transition zone. If atomic diffusion is efficient, details of the initial shape of the profile become less important in the course of evolution. Diffusion broadens the chemical gradients, and results in less effective mode trapping and different pulsation periods. Furthermore, we report on the possible consequences of the He-flash for the $g$-modes. The outer edge of a flash-induced convective region introduces an additional chemical transition in the stellar models, and the corresponding spike in the Br\\\"unt-V\\\"ais\\\"al\\\"a frequency produces a complicated mode trapping signature in the period spacings. } {} ", "introduction": "Hot B-type subdwarfs are identified as core He-burning stars surrounded by a very thin H-envelope \\citep{heber1986}. This places them at the blue extension of the horizontal branch in the Hertzsprung-Russell diagram. Hence, they are also referred to as extreme horizontal branch (EHB) stars. Subdwarf B (sdB) stars are ubiquitous in our Galaxy, where they dominate the population of faint blue objects at high galactic latitudes \\citep{green1986}. They are also believed to be responsible for the ultraviolet excess or `UV upturn' in giant elliptical galaxies \\citep{brown1997,yi1997}. A thorough review of hot subdwarfs is given by \\citet{heber2009}. The future evolution of sdB stars is straightforward and undisputed. After He in the core is exhausted, a short phase of He-shell burning follows, and the star may be identified as a hotter subdwarf of O-type (sdO). The H-envelope is too thin to sustain H-shell burning, so the star will end as a C-O white dwarf without evolving through the asymptotic giant branch phase. The formation of sdBs on the other hand, remains a much debated topic. Many formation channels have been proposed such as enhanced mass loss on the RGB \\citep{d'cruz1996}, mass loss through binary interaction \\citep{mengel1976}, and the mergers of two He white dwarfs \\citep{webbink1984}. The relative importance of different formation scenarios has been evaluated by binary population synthesis studies \\citep{han2002,han2003}. While these studies are valuable, it should not be forgotten that they are dependent on parametric descriptions of binary interaction. Detailed mass determinations of subdwarf B stars could provide important constraints on the binary interaction mechanism. Accurate astrophysical mass determinations are only possible in special circumstances such as in eclipsing binary systems. Asteroseismology provides an alternative method, since it allows a detailed study of the interior of non-radially pulsating stars. The consistency of both approaches has been achieved for \\object{PG\\,1336$-$018} (see \\citealt{vuckovic2007}, \\citealt{hu2007}, and \\citealt{charpinet2008}). Subdwarf B stars exhibit a variety of pulsations. The first sdB pulsator, \\object{EC\\,14026$-$2647}, was observed by \\citet{kilkenny1997} to pulsate in multiple short-period modes. This prototype represents the variable class \\object{V361\\,Hya} stars which now includes 42 rapid sdB pulsators, with periods in the range \\mbox{80-600 s}. The short-period modes have been interpreted as low radial order, low spherical degree $p$-modes \\citep{charpinet1997}. The \\object{V361\\,Hya} stars are amongst the hottest sdB stars with effective temperatures between 28 000 K and 35 000 K and surface gravities $5.2<\\log g<6.1$. At the cooler end of the EHB, 31 sdB pulsators with periods between 30 min and two hours have been discovered by \\citet{green2003}. This variable class is named V1093 Her but is also often referred to as \\object{PG\\,1716$+$426} stars after the class prototype. The long periods suggest that these stars pulsate in high radial order $g$-modes. Especially interesting are the sdB pulsators that exhibit both $p$- and $g$-mode pulsations. Three of these so-called hybrid pulsators have been found at the intersection of the \\object{V361\\,Hya} and \\object{V1093\\,Her} stars \\citep{oreiro2005,baran2005,schuh2006,lutz2008}. The opacity mechanism operating in the Fe opacity bump around $\\log T=5.3$ has been successfully used to explain the excitation of $p$- as well as the $g$-modes \\citep{charpinet1996,fontaine2003}. This opacity bump is caused by Fe accumulation owing to the competing diffusion processes of gravitational settling and radiative levitation. While seismic mass determinations have been achieved for a dozen \\object{V361\\,Hya} stars using static envelope models (see \\citealt{fontaine2008} and references therein), this is not yet the case for \\object{V1093\\,Her} stars. The reason is twofold. First of all, the $g$-modes have lower amplitudes and longer pulsation periods and thus require higher precision observations and longer observing runs to detect the frequencies. Secondly, envelope models may suffice to describe the \\object{V361\\,Hya} stars' shallow $p$-modes that probe only the outer layers. However, they are not suitable for modelling $g$-modes that penetrate to deeper stellar regions. Stellar models with detailed information about the He-rich core are required to predict reliable $g$-mode periods. Another problem for the V1093 Her stars is the discrepancy between the observed and theoretically predicted instability strip. The first theoretical models that show unstable $g$-mode pulsations \\citep{fontaine2003} have much lower effective temperatures than observed with a discrepancy of $\\sim$ 5000 K. \\citet{jeffery2006b} managed to place the blue edge of the instability strip within $\\sim$1000 K of the observed blue edge. They did this by using OP opacities (\\citealt{badnell2005}) rather than OPAL opacities (\\citealt{iglesias1996}), and by enhancing the Ni abundance in the envelope in addition to Fe. Thus, their work suggests that sophisticated models including up-to-date opacities and diffusion of other Fe-group elements are required to address the instability issue properly. An exploratory study of how the pulsational properties of sdB stars are related to their internal structure has been carried out by \\citet{charpinet2000}. They showed that the He-H transition zone between the He-rich core and H-rich envelope is responsible for trapping and confining $g$-modes in sdB stars. This is similar to the well-known mode trapping phenomenon in compositionally stratified white dwarfs \\citep{winget1981}. Mode trapping results in deviations from the asymptotic constant period spacing, making the $g$-modes sensitive to the mass of the H-envelope. Furthermore, the C-O/He-transition region between the convective and the radiative part of the core will also influence the $g$-mode period distribution. This may be a new interesting way to follow the He-core evolution, which would also have implications in a broader context for horizontal branch stars. Here, we shall explore the sensitivity of the $g$-modes to certain assumptions about the He-flash. This energetic event characterizes the start of He-ignition in degenerate cores of low-mass stars ($<$2 M$_{\\odot}$). Most stellar evolution codes encounter numerical difficulties in calculating this phase, and even in successful cases uncertainties remain because of the approximative treatment of convection. Three-dimensional hydrodynamic calculations are necessary to investigate this enigmatic phase of stellar evolution (e.g., \\citealt{deupree1996,dearborn2006,mocak2008}), but these are expensive in computing time. When many models are needed, as in forward modelling, it is more practical to construct post-flash models without following the previous evolution in detail. One should then use the outcome of detailed studies to ensure that the stellar models are as realistic as possible. Most simulations indicate that the initial flash starts off-centre, followed by several mini-flashes moving towards the centre. While it is generally found that most of the inner region will be convectively mixed, there are uncertainties in the outer extent of the convective region and the amount of He-burning during the He-flash varying from 3\\% to 7\\% in mass fraction \\citep{piersanti2004,serenelli2005}. The edge of the He-flash-induced convective region will leave a chemical composition gradient, which we believe can produce additional mode trapping features. We start with constructing zero age extreme horizontal branch (ZAEHB) models in Sect.~\\ref{zaehb}. We make an analytical fit to the H-profile in the core-envelope transition region and examine its influence on the $g$-mode periods (Sect.~\\ref{profile}). The effects of uncertainties in the He-flash treatment, i.e.~convective mixing and C-production, are explored in Sect.~\\ref{heflash}. In Sect.~\\ref{ehb}, we perform a stability analysis on a grid of evolutionary sdB structures during the core He-burning phase. We show the impact of the opacities and H/He-diffusion on the instability strip. The conclusions are summarized in Sect.~\\ref{conclusions}. ", "conclusions": "The precise details of chemical transitions in subdwarf B stars have a considerable influence on the period spacings of high order $g$-modes. This is a signature of mode trapping that occurs due to discontinuities in the BV-frequency. We have shown that the structure of mode trapping depends sensitively on the exact shape of the H-profile in the core-envelope transition zone. Thus, when constructing ZAEHB models without following the previous evolution, the chemical profiles must be modelled carefully. Since the action of mode trapping is so sensitive to the shape of the chemical transition, it is necessary to include atomic diffusion during the evolution, i.e., the processes of gravitational settling, temperature and concentration diffusion. Furthermore, if the sdB star started He-fusion in a run-away flash, the inner part of the core is convectively mixed and C is produced up to $\\sim$7\\%. The edge of the He-flash-induced convective core is accompanied by a chemical transition. We showed that this can cause additional mode trapping features resulting in a complicated behaviour of the period spacings. Multiple composition gradients lead to simultaneous mode trapping in different regions. The $g$-mode spectra become very complicated, and it will be difficult to directly infer detailed information about the composition gradients from observed $g$-modes. It is therefore crucial to evaluate the importance of different physical processes, and to account for all of them realistically, as we do here. Finally, and perhaps most importantly, we report on a possible step forward in solving the `blue edge problem' of the V1093 Her instability strip. Our evolutionary sdB models with (i) OP opacities, (ii) the inclusion of H/He diffusion, and (iii) a parametric Fe/Ni enhancement, show unstable $l\\leq 3$ $g$-modes for effective temperatures up to $30,000$ K. This is only achieved when all three ingredients are included in the computations. It should be noted that it remains unclear which $l$-values the observed long-period pulsation modes have. Thus, although we have demonstrated the importance of H/He diffusion, solving the blue edge problem effectively will require mode identification, in addition to a realistic treatment of radiative levitation and other transport mechanisms." }, "0910/0910.3701_arXiv.txt": { "abstract": "Considering the high-energy limit of the QCD gluon distribution inside a nucleus, we calculate the proton-nucleus total inelastic cross section using a simplified dipole model. We show that, if gluon saturation occurs in the nuclear surface region, the total cross section of proton-nucleus collisions increases more rapidly as a function of the incident energy compared to that of a Glauber-type estimate. We discuss the implications of this with respect to recent ultra-high-energy cosmic ray experiments. ", "introduction": "In the ultra-high-energy domain, the mechanism of inelastic hadronic collisions is dominated by the contribution from small-$x$ gluons. This is the basic reason why the total or inelastic proton-proton collision cross section increases as a function of the incident energy. A simple way to see this is to use the eikonal expression for the reaction cross section in the impact parameter representation as% \\begin{equation} \\sigma _{r}=\\int d^{2}\\mathbf{b}[1-\\exp (-2\\chi (\\mathbf{b},s))], \\label{inela} \\end{equation}% where the eikonal $\\chi (\\mathbf{b},s)$ counts essentially the total number of possible scattering centers of the constituents inside the target \\textquotedblleft seen\\textquotedblright\\ by the projectile passing through the target along a straight line with the impact parameter $\\mathbf{b,}$ and with center-of-mass energy $\\sqrt{s}$. If the target is thick enough, the eikonal function is much larger than unity ( $\\chi (\\mathbf{b},s)\\gg 1$) in the central region, but it falls down to zero in the surface region. The integrand of Eq.(\\ref{inela}) keeps a value almost equal to unity while $\\chi (\\mathbf{b},s)\\gg 1,$ and falls down quickly to zero near the surface. From this we may determine an effective radius $b_{1/2}$ such that the integral (\\ref{inela}) is approximately given by \\begin{equation} \\sigma _{r}\\simeq \\pi b_{1/2}^{2}. \\end{equation}% Here $b_{1/2}$ might be estimated by fixing the value of $\\chi (\\mathbf{b}% _{1/2},s)$, for example, as \\begin{equation} \\chi (\\mathbf{b}_{1/2},s)=\\ln 2, \\end{equation}% so that the effective radius depends on the energy $b_{1/2}=b_{1/2}\\left( \\sqrt{s}\\right) $. Let us assume that the eikonal function factorizes in the form \\begin{equation} \\chi (\\mathbf{b},s)=P(\\mathbf{b})N(\\sqrt{s}), \\end{equation}% where $P(\\mathbf{b})$ is the probability distribution of the scattering center (e.g., partons) in the transverse plane and $N(x)$ is the number of partons which can interact for a given $\\sqrt{s}$. Suppose that $P(\\mathbf{b}% )$ is a two-dimensional Gaussian distribution of width $R,$ \\begin{equation} P(\\mathbf{b})=\\frac{1}{\\left( \\pi R\\right) ^{2}}e^{-\\frac{b^{2}}{R^{2}}} \\end{equation}% and that, for large $\\sqrt{s}$, the number of partons increase with $\\sqrt{s} $ as \\begin{equation} N(x)\\propto \\sqrt{s}^{\\alpha }. \\end{equation}% Then, we have \\begin{equation} b_{1/2}^{2}\\left( \\sqrt{s}\\right) =\\alpha R^{2}\\ln \\sqrt{s}+Const. \\end{equation}% In such a situation, the reaction cross section increases as function of the incident energy for very large $\\sqrt{s}$ as \\begin{equation} \\sigma _{r}\\simeq \\alpha \\pi R^{2}\\ln \\sqrt{s}. \\label{sig1} \\end{equation} If the edge of the distribution has an exponential tail instead of a Gaussian distribution, a similar argument will show that the cross section would increase as \\begin{equation} \\sigma _{r}\\simeq Const\\times \\left( \\ln \\sqrt{s}\\right) ^{2}. \\label{sig2} \\end{equation} The important point of this simple argument is that the rate of increase is related to the diffuseness $R$ of the probability distribution of the scattering center. The more diffuse the surface thickness is, the more quickly the reaction cross section increases as function of the incident energy. The possibility of such a mechanism, not only in the proton-proton but also in nucleus-nucleus cross section, was suggested many years ago~\\cite% {Kodama,Kodama1}. In this paper we explore the idea of~\\cite{Kodama,Kodama1} in the language of the QCD gluon saturation mechanism for the proton-nucleus reaction. We show that, if the gluon distribution becomes saturated at some energy scale inside the nuclear surface region, then the reaction cross section of proton-nucleus collisions starts to increase very quickly and eventually overcomes the values estimated by the usual Glauber type of calculation. Applying a simple effective dipole model for the reaction mechanism, we find that such an energy scale is of the order of $10^{17}-10^{18}eV.$ Above this energy scale, the behavior of the proton-nucleus cross section begins to change. We suggest that such an energy dependence of the proton-nucleus cross section may be observed in terms of the quantity called $\\left\\langle X_{\\max }\\right\\rangle $ in the air showers of ultra-high-energy cosmic rays. Using a very simple toy model estimate of $\\left\\langle X_{\\max }\\right\\rangle $, we show that our calculated values of the proton-nucleus reaction cross section are consistent with the recently observed $% \\left\\langle X_{\\max }\\right\\rangle $ by the Pierre Auger Observatory experiments~\\cite{Auger_Xmax} for incident protons at ultra-high energies. ", "conclusions": "In this paper, we have explored the idea of gluon saturation inside a nucleus in the high-energy limit and its effect on the proton-nucleus cross section. We show that if gluon saturation occurs in the surface region of the target nucleus, the proton-nucleus cross section starts to increase very rapidly as a function of the incident energy. Such a mechanism should eventually happen for an ultra-high energy scale, but the question is in which energy scale such a scenario starts to dominate. Such energy scale is determined by the form of the distribution of gluons near the surface area. If we assume that the small $x$ gluon distribution in the nucleus follows that of the nucleon wave function inside the nucleus, the energy scale where the gluon saturation scenario starts to dominate the independent nucleon picture is around $10^{17}-10^{18}$ eV. We note that different profile functions give more or less the same energy scale once the proton-proton cross section is well fitted. It is very suggestive that the gluon saturation scenario inside the surface area seems consistent with the proton primary interpretation of the observed $\\left\\langle X_{\\max }\\right\\rangle $ behavior in the Pierre Auger Laboratory experiment. As we see, the difference between the two pictures, INGP and GSNS, becomes very large at high energies. The reason for this is that, while in the INGP the effect of virtual gluons which bound the nucleon near the surface area is completely neglected, these gluons become dominant at high energies in the GSNS scenario. In a simple-minded argument, one might think that such an effect of nuclear binding must be negligible at high energies, since the ratio of the binding energy of a nucleon to the incident energy tends to zero. However, the situation may not be so simple. In the GSNS scenario, the density of virtual gluons, probably forming a kind of fractal fingers when penetrating into the vacuum among nucleons similar to the electric discharge pattern, becomes higher and higher at high energies, and eventually percolate everywhere even in the nuclear surface region. According to the color glass condensate picture~\\cite{Larry,Larry2,Larry3,Larry4}, such a scenario should happen at some energy scale, even at the lowest density region of the nuclear surface. Naturally, the energy scale depends on the precise form of the geometric gluon distribution inside the nucleus. If the distribution does not follow the probability distribution of nucleons but more sharp surface distribution, then the energy scale may shift to higher energy. In the example of Fig. 3 and Fig. 4 the surface thickness parameter of the Woods-Saxon distribution is taken a little bit smaller than the usual value fitted to the nuclear distribution in nucleus ($d\\approx 0.6~fm$) since this fit only applies for heavy nuclei ($A>40$). If we take $d=0.6~fm,$ the energy scale is lower by one order of magnitude. Therefore, the energy scale depends crucially on how the gluons starts to saturate in the nuclear surface region. Depending on this, the energy scale can be even lower than estimated here. To find a real energy scale where the gluon saturation occurs at the nuclear surface, further investigation on high-energy proton-nucleus or electron-nucleus collisions will be necessary. There exist ambiguities in the proton-proton cross section used in this paper (those of the model and extrapolation of the PDF, as well as those of experimental data) and these affect the precise value of the energy scale. However, the general conclusion of the present work does not change, as far as the energy dependence of the proton-proton cross section is fixed." }, "0910/0910.4910_arXiv.txt": { "abstract": "{The identification and study of the first galaxies remains one of the most exciting topics in observational cosmology. The determination of the best possible observing strategies is a very important choice in order to build up a representative sample of spectroscopically confirmed sources at high-z ($z\\gtapprox7$), beyond the limits of present-day observations. } {This paper is intended to precisely adress the relative efficiency of lensing and blank fields in the identification and study of galaxies at $6\\ltapprox z \\ltapprox 12$. } {The detection efficiency and field-to-field variance are estimated from direct simulations of both blank and lensing fields observations. Present known luminosity functions in the UV are used to determine the expected distribution and properties of distant samples at $z\\gtapprox6$ for a variety of survey configurations. Different models for well known lensing clusters are used to simulate in details the magnification and dilution effects on the backgound distant population of galaxies. } {The presence of a strong-lensing cluster along the line of sight has a dramatic effect on the number of observed sources, with a positive magnification bias in typical ground-based ``shallow'' surveys ($AB\\ltapprox25.5$). The positive magnification bias increases with the redshift of sources and decreases with both depth of the survey and the size of the surveyed area. The maximum efficiency is reached for lensing clusters at $z\\sim0.1-0.3$. Observing blank fields in shallow surveys is particularly inefficient as compared to lensing fields if the UV LF for LBGs is strongly evolving at $z\\gtapprox7$. Also in this case, the number of $z\\ge8$ sources expected at the typical depth of JWST ($AB\\sim28-29$) is much higher in lensing than in blank fields (e.g. a factor of $\\sim10$ for $AB\\ltapprox28$). {All these results have been obtained assuming that number counts derived in clusters are not dominated by sources below the limiting surface brightness of observations, which in turn depends on the reliability of the usual scalings applied to the size of high-z sources. } } {Blank field surveys with a large field of view are needed to prove the bright end of the LF at $z\\gtapprox6-7$, whereas lensing clusters are particularly useful for exploring the mid to faint end of the LF. } ", "introduction": " ", "conclusions": "Summary and conclusions} We have evaluated the relative efficiency of lensing clusters with respect to blank fields in the identification and study of $z\\ge6$ galaxies. The main conclusions of this study are given below. For magnitude-limited samples of LBGs at $z\\ge6$, the magnification bias increases with the redshift of sources and decreases with both the depth of the survey and the size of the surveyed area. Given the typical near-IR FOV in lensing fields, the maximum efficiency is reached for clusters at $z\\sim0.1-0.3$, with maximum cluster-to-cluster differences ranging between 30 and 50\\% in number counts, depending on the redshift of sources and the LF. The relative efficiency of lensing with respect to blank fields strongly depends on the shape of the LF, for a given photometric depth and FOV. The comparison between lensing and blank field number counts is likely to yield strong constraints on the LF. The presence of a strong-lensing cluster along the line of sight has a dramatic effect on the observed number of sources, with a positive magnification effect in typical ground-based ``shallow'' surveys ($AB\\le25.5$). The postive magnification bias increases with the redshift of sources, and also from optimistic to pessimistic values of the LF. In case of a strongly evolving LF at $z\\ge7$, as proposed by \\cite{Bouwens08}, blank fields are particularly inefficient as compared to lensing fields. For instance, the size of the surveyed area in ground-based observations would need to increase by a factor of $\\sim10$ in blank fields with respect to a typical $\\sim30-40$ $arcmin^{2}$ survey in a lensing field, in order to reach the same number of detections at $z\\sim6-8$, and this merit factor increases with redshift. All these results have been obtained assuming that number counts derived in clusters are not dominated by sources below the limiting surface brightness of observations, which in turn depends on the reliability of the usual scalings applied to the size of high-z sources. Ground-based ``shallow'' surveys are dominated by field-to-field variance reaching $\\sim$ 30 to 50\\% in number counts between z$\\sim$6 and 8 in a unique $\\sim30-40$ $arcmin^{2}$ lensing field survey (or in a 400 $arcmin^{2}$ blank field), assuming a strongly evolving LF. The number of z$>$8 sources expected at the typical depth of JWST ($AB\\sim28-29$) is much higher in lensing than in blank fields if the UV LF is rapidly evolving with redshift (i.e. a factor of $\\sim10$ at $z\\sim10$ with $AB\\ltapprox28$). Blank field surveys with a large FOV are needed to probe the bright edge of the LF at $z\\ge6-7$, whereas lensing clusters are particularly useful to explore the mid to faint end of the LF." }, "0910/0910.4584_arXiv.txt": { "abstract": "{We determine the solar neutrino fluxes from a global analysis of the solar and terrestrial neutrino data in the framework of three-neutrino mixing. Using a Bayesian approach we reconstruct the posterior probability distribution function for the eight normalization parameters of the solar neutrino fluxes plus the relevant masses and mixing, with and without imposing the luminosity constraint. This is done by means of a Markov Chain Monte Carlo employing the Metropolis-Hastings algorithm. We also describe how these results can be applied to test the predictions of the Standard Solar Models. Our results show that, at present, both models with low and high metallicity can describe the data with good statistical agreement.} \\preprint{YITP-SB-09-34\\\\IFT-UAM/CSIC-09-50\\\\EURONU-WP6-09-11} ", "introduction": "The idea that the Sun generates power through nuclear fusion in its core was first suggested in 1919 by Sir Arthur Eddington who pointed out that the nuclear energy stored in the Sun could explain the apparent age of the Solar System. In 1939, Hans Bethe described in an epochal paper~\\cite{Bethe:1939bt} two nuclear fusion mechanisms by which main sequence stars like the Sun could produce the energy necessary to power their observed luminosities. The two mechanisms have become known as the pp-chain and the CNO-cycle~\\cite{Bahcall:1989ks}. For both chains the basic energy source is the burning of four protons to form an alpha particle, two positrons, and two neutrinos. In the pp-chain, fusion reactions among elements lighter than $A = 8$ produce a characteristic set of neutrino fluxes, whose spectral energy shapes are known but whose fluxes must be calculated with a detailed solar model. In the CNO-cycle the abundance of \\Nuc[12]{C} plus \\Nuc[13]{N} acts as a catalyst, while the \\Nuc[13]{N} and \\Nuc[15]{O} beta decays provide the primary source of neutrinos. In order to precisely determine the rates of the different reactions in the two chains, which are responsible for the final neutrino fluxes and their energy spectrum, a detailed knowledge of the Sun and its evolution is needed. Standard Solar Models (SSM's)~\\cite{Bahcall:1987jc, TurckChieze:1988tj, Bahcall:1992hn, Bahcall:1995bt, Bahcall:2000nu, Bahcall:2004pz, PenaGaray:2008qe} describe the properties of the Sun and its evolution after entering the main sequence. The models are based on a set of observational parameters (the present surface abundances of heavy elements and surface luminosity of the Sun, as well as its age, radius and mass) and on several basic assumptions: spherical symmetry, hydrostatic and thermal equilibrium, and equation of state. Over the past five decades the solar models were steadily refined as the result of increased observational and experimental information about the input parameters (such as nuclear reaction rates and the surface abundances of different elements), more accurate calculations of constituent quantities (such as radiative opacity and equation of state), the inclusion of new physical effects (such as element diffusion) and the development of faster computers and more precise stellar evolution codes. Despite the progress of the theory, only neutrinos, with their extremely small interaction cross sections, can enable us to see into the interior of a star and thus verify directly our understanding of the Sun~\\cite{Bahcall:1964gx}. Indeed from the earliest days of solar neutrino research this test has been a primary goal of solar neutrino experiments, but for many years the task was made difficult by the increasing discrepancy between the predictions of the SSM's and the solar neutrino observations. This so-called ``solar neutrino problem''~\\cite{Bahcall:1968hc, Bahcall:1976zz} was finally solved by the modification of the Standard Model of Particle Physics with the inclusion of neutrino masses and mixing. In this new framework leptonic flavors are no longer symmetries of Nature, and neutrinos can change their flavor from the production point in the Sun to their detection on the Earth. This flavor transition probability is energy dependent~\\cite{Pontecorvo:1967fh, Gribov:1968kq, Wolfenstein:1977ue, Mikheev:1986gs}, which explains the apparent disagreement among experiments with different energy windows. This mechanism is known as the LMA-MSW solution to the solar neutrino problem, and affects both the overall number of events in solar neutrino experiments and the relative contribution expected from the different components of the solar neutrino spectrum. Due to these complications, at first it was necessary to assume the SSM predictions for all the solar neutrino fluxes and their uncertainties in order to extract reasonably constrained values for neutrino masses and mixing. The upcoming of the real-time experiments Super-Kamiokande and SNO and the independent determination of the flavor oscillation probability using reactor antineutrinos at KamLAND opened up the possibility of extracting the solar neutrino fluxes and their uncertainties directly from the data~\\cite{Garzelli:2001ju, Bahcall:2002zh, Bahcall:2002jt, Bahcall:2004ut, Bahcall:2003ce, Bandyopadhyay:2006jn, PenaGaray:2008qe}. Nevertheless, in these works some set of simplifying assumptions had to be imposed in order to reduce the number of free parameters to be determined. In parallel to the increased precision of the SSM-independent determination of the neutrino flavor parameters, a new puzzle has emerged in the consistency of SSM's~\\cite{Bahcall:2004yr}. Till recently SSM's have had notable successes in predicting other observations. In particular, quantities measured by helioseismology such as the radial distributions of sound speed and density~\\cite{Bahcall:1992hn, Bahcall:1995bt, Bahcall:2000nu, Bahcall:2004pz} showed good agreement with the predictions of the SSM calculations and provided accurate information on the solar interior. A key element to this agreement is the input value of the abundances of heavy elements on the surface of the Sun~\\cite{Grevesse:1998bj}. However, recent determinations of these abundances point towards substantially lower values than previously expected~\\cite{Asplund:2004eu, Asplund:2009fu}. A SSM which incorporates such lower metallicities fails at explaining the helioseismological observations~\\cite{Bahcall:2004yr}, and changes in the Sun modeling (in particular of the less known convective zone) are not able to account for this discrepancy~\\cite{Chaplin:2007uh, Basu:2006vh}. So far there has not been a successful solution of this puzzle. Thus the situation is that, at present, there is no fully consistent SSM. This led to the construction of two different sets of SSM's, one (labeled ``GS'') based on the older solar abundances~\\cite{Grevesse:1998bj} implying high metallicity, and one (labeled ``AGS'') assuming lower metallicity as inferred from more recent determinations of the solar abundances~\\cite{Asplund:2004eu, Asplund:2009fu}. In Ref.~\\cite{PenaGaray:2008qe} the solar fluxes corresponding to such two models were detailed, based on updated versions of the solar model calculations presented in Ref.~\\cite{Bahcall:2004pz}. These fluxes were denoted as ``BPS08(GS)'' and ``BPS08(AGS)'', respectively. In a very recent work~\\cite{Serenelli:2009yc} an update of the BPS08(AGS) solar model has been constructed using the latest determination of the compositions~\\cite{Asplund:2009fu} as well as some improvement in the equation of state. For what concerns the overall normalization of solar neutrino fluxes, the predictions of this new model are very close to those of BPS08(AGS). In this work we perform a solar model independent analysis of the solar and terrestrial neutrino data in the framework of three-neutrino masses and mixing. The aim of this analysis is to simultaneously determine the flavor parameters and all the solar neutrino fluxes with a minimum set of theoretical priors. In Sec.~\\ref{sec:ana} we present the method employed, the data included in the analysis and the physical assumptions used in this study. The results of the analysis are given in Sec.~\\ref{sec:res}, where we show the reconstructed posterior probability distribution function for the eight normalization parameters of the solar neutrino fluxes. We discuss in detail the effect of the luminosity constraint~\\cite{Bahcall:2001pf} as well as the role of the Borexino experiment and its potential for improvement. In addition, we use the results of this analysis to statistically test to what degree the present solar neutrino data can discriminate between the two SSM's. Finally in Sec.~\\ref{sec:sum} we summarize our conclusions. ", "conclusions": "\\label{sec:sum} We have performed a solar model independent analysis of the solar and terrestrial neutrino data in the framework of three-neutrino oscillations, following a Bayesian approach in terms of a Markov Chain Monte Carlo using the Metropolis-Hastings algorithm. This approach has allowed us to reconstruct the probability distribution function in the entire eleven-dimensional parameter space, consistently incorporating the required set of theoretical priors. The best fit values and allowed ranges for the eight solar neutrino fluxes are summarized in Eq.~\\eqref{eq:bestlc} and Fig.~\\ref{fig:fitlc} for the analysis with the luminosity constraint, and in Eq.~\\eqref{eq:bestnolc} and Fig.~\\ref{fig:fitnolc} for the more general case of unconstrained solar luminosity. We found that at present the neutrino-inferred luminosity perfectly agrees with the measured one and it is known with a $1\\sigma$ uncertainty of 14\\%. We have also tested the fractional energy production in the pp-chain and the CNO-cycle, finding that the total amount of the solar luminosity produced in the CNO-cycle is bounded to be $L_\\text{CNO} / L_\\odot < 2.8\\%$ at 99\\% CL irrespective of whether the luminosity constraint is imposed or not. We have then presented a statistical test which can be performed with these results in order to shed some light on the so-called solar composition problem, which at present arises in the construction of the Standard Solar Model. We found that the low value of the \\Nuc[8]{B} flux measured at SK and SNO points towards low metallicity models, whereas the measurement of \\Nuc[7]{Be} in Borexino and the corresponding value of the \\Nuc{pp} flux implied by the luminosity constraint show better agreement with high metallicity models. Altogether the fit shows a slight preference for models with higher metallicities, however the difference between the two models is not very significant at statistical level. While a realistic improvement of the Borexino data analysis in the near future can positively affect the direct determination of most solar neutrino fluxes, it is unlikely that enough precision will be achieved to go beyond the the present theoretical uncertainties of the SSM's. The largest difference between the models lies on the CNO fluxes that give predictions which differ by about 30\\%. Thus ideally in order to achieve a statistically meaningful discrimination between the models one would need a low energy solar neutrino experiment capable of measuring the neutrino energy spectrum for energies between $0.5~\\text{MeV} \\lesssim E_\\nu \\lesssim 1.5~\\text{MeV}$ and, more importantly, of rejecting the radioactive backgrounds to the required level, so to allow for a determination of the CNO fluxes at $\\mathcal{O}(\\text{30\\%})$ precision." }, "0910/0910.4482.txt": { "abstract": "We present high precision radial velocities (RVs) of double-lined spectroscopic binary stars HD78418, HD123999, HD160922, HD200077 and HD210027. They were obtained based on the high resolution echelle spectra collected with the Keck I/Hires, Shane/CAT/Hamspec and TNG/Sarge telescopes/spectrographs over the years 2003-2008 as a part of TATOOINE search for circumbinary planets. The RVs were computed using our novel iodine cell technique for double-line binary stars which relies on tomographically disentangled spectra of the components of the binaries. The precision of the RVs is of the order of 1-10 m~s$^{-1}$ and to properly model such measurements one needs to account for the light time effect within the binary's orbit, the relativistic effects and the RV variations due to the tidal distortions of the components of the binaries. With such proper modeling, our RVs combined with the archival visibility measurements from the Palomar Testbed Interferometer allow us to derive very precise spectroscopic/astrometric orbital and physical parameters of the binaries. In particular, we derive the masses, the absolute K and H band magnitudes and the parallaxes. The masses together with the absolute magnitudes in the K and H bands enable us to estimate the ages of the binaries. These RVs allow us to obtain some of the most accurate mass determinations of binary stars. The fractional accuracy in $m\\sin i$ only and hence based on the RVs alone ranges from 0.02\\% to 0.42\\%. When combined with the PTI astrometry, the fractional accuracy in the masses ranges in the three best cases from 0.06\\% to 0.5\\%. Among them, the masses of HD210027 components rival in precision the mass determination of the components of the relativistic double pulsar system PSR~J0737-3039. In the near future, for double-lined eclipsing binary stars we expect to derive masses with a fractional accuracy of the order of up to $\\sim$0.001\\% with our technique. This level of precision is an order of magnitude higher than of the most accurate mass determination for a body outside the Solar System --- the double neutron star system PSR~B1913+16. ", "introduction": "The first observations of a spectroscopic binary star, $\\zeta$ UMa (Mizar), were announced by Edward C. Pickering (1846-1919) on 13 November 1889 during a meeting of the National Academy of Sciences in Philadelphia \\citep{Pickering:90::}. A similar announcement about $\\beta$ Per (Algol) was made by Herman C. Vogel (1841-1907) on 28 November 1889 during a session of Konglich-Preussiche Akademie der Wissenschaften \\citep{Vogel:90a::,Vogel:90b::}. Even though a transatlantic telegraph cable had been available since the 1860s, it was quite unusual timing for a pre-astro-ph era. Pickering noted that the K line of Mizar occasionally appeared double and this way discovered the first double-lined spectroscopic binary star. Vogel measured radial velocities (RVs) of Algol and used them to prove that the known variations in the brightness of Algol are indeed caused by a ``dark satellite revolving about it\" \\citep{Vogel:90b::}. Vogel and collaborators built a series of prism spectrographs for the 30-cm refractor of the Potsdam Astrophysical Observatory \\citep{Vogel:00::}. In 1888, they initiated a photographic radial-velocity program. Vogel was not the first to take a photograph of a stellar spectrum but improved the technique and obtained an RV precision of 2-4 km~s$^{-1}$ \\citep{Vogel:91::}. Vogel (1906) and Pickering (1908) were both awarded the Bruce medal. Yet another Bruce medalist (1915), William W. Campbell (1862-1938) and collaborators carried out a large spectroscopic program at the Lick Observatory and discovered many spectroscopic binaries \\citep{Camp:11::}. They prepared the first catalog of spectroscopic binaries \\citep{Camp:05::}. Vogel was the director of the Potsdam Observatory for 25 years, Pickering of the Harvard College Observatory for 42 years and Campbell of the Lick Observatory for 30 years. The administrative duties must have been not too distracting those days. Both Vogel and Campbell recognized the importance of flexure and temperature control of a spectrograph \\citep{Vogel:90c::,Camp:98::,Camp:00::} and Campbell also noted the issue of a slit illumination and its impact on RV precision \\citep{Camp:16::}. Today a high RV precision is achieved by either using a very stable fiber fed spectrograph that is contained in a controlled environment (a vacuum tank) or using an absorption cell to superimpose a reference spectrum onto a stellar spectrum and this way measure and account for the systematic RV errors. Current state of the art precision is at the level of $\\sim$1 m~s$^{-1}$. It is however important to note that such a precision refers to single stars or at best single-lined spectroscopic binaries where the influence of the secondary spectrum can be neglected. In such a case, given a stable spectrograph, an RV measurement is essentially a measurement of a shift of an otherwise constant shape (spectrum). Radial velocities (RVs) of double-lined spectroscopic binary stars (SB2) can be used effectively to derive basic parameters of stars if the stars happen to be eclipsing or their astrometric relative orbit can be determined. It is quite surprising that the RV precision of double-lined binary stars on the average has not improved much over the last 100 years (see Fig.~\\ref{fig1}). With the exception of our previous work \\citep{Konacki:05::,Konacki:09a::}, the RV precision for such targets typically varies from $\\sim$0.1 km~s$^{-1}$ to $\\sim$1 km~s$^{-1}$ and clearly is much worse than what has been achieved for stars with planets or single-lined binary stars. The main problem with double-lined binary stars is that one has to deal with two sets of superimposed spectral lines whose corresponding radial velocities change considerably with typical amplitudes of $\\sim$ 50-100 km~s$^{-1}$. In consequence a spectrum is highly variable and obviously one cannot measure RVs by noting a simple shift. We have developed a novel iodine cell based approach that employs a tomographic disentangling of the component spectra of SB2s and allows one to measure RVs of the components of SB2s with a precision of the order 1-10 m$\\,$s$^{-1}$ \\citep{Konacki:09a::,Konacki:09b::}. Such quality RVs not only enable us to search for circumbinary extrasolar planets \\citep{Konacki:09b::} but also to determine basic parameters of stars with an unprecedented precision. In particular the masses of stars for non eclipsing SB2s can be easily determined with a fractional accuracy of the order of at least $\\sim$$0.1\\%$ and often even $\\sim$$0.01\\%$. Moreover, we expect that the accuracy in masses will reach $\\sim$$0.001\\%$ level when our method is applied to eclipsing binary stars. Such a level of precision is an order of magnitude higher than of the most accurate mass determination for a body outside the Solar System --- the double neutron star system PSR~B1913+16 \\citep{Nice:08::}. Below we present our precision RV data sets for five targets HD78418, HD123999, HD160922, HD200077 and HD210027 from our on-going TATOOINE (The Attempt To Observe Outer-planets In Non-single-stellar Environments) RV program to search for circumbinary planets. All of them have been extensively observed with the Palomar Testbed Interferometer \\citep[PTI;]{Col:99a::}. The archival PTI visibility measurements can be used to derive relative astrometric orbits of the binaries. These combined with our spectroscopic orbits allow for a complete orbital and physical description of the systems (with the exception of the radii of the components). In \\S{2} we describe the RV measurements and their modeling and in \\S{3} the visibility measurements and their modeling. In \\S{4} we present the spectroscopic and astrometric orbital solutions and the resulting orbital and physical parameters of the binaries. A discussion is provided in \\S{5}. ", "conclusions": "There are a few ways to determine accurate masses of normal stars. Perhaps the most classic one is through absolute astrometric orbits of both components of a binary. However this constitues a challenging measurement and in practice masses are typically derived using diverse data sets e.g. (1) by combining RVs, relative astrometry and parallax for single-lined spectroscopic binaries, (2) as in this paper by combining relative astrometry and RVs for double-lined spectroscopic binaries and (3) by combining light curves and RVs for eclipsing double-lined binaries. The last method is the most useful one as it not only provides the most accurate masses due to a convenient edge-on geometry (note that the masses are derived from their respective M $\\sin^3i$) but also enables one to determine the radii of stars. In a recent review, \\cite{Torres:09::} collect 118 detached binary stars (including 94 eclipsing) with the most accurate mass determinations in the literature. These are denoted with open circles in Fig.~18. Even though our targets are not eclipsing and the orbital inclinations and their errors have a significant impact on the precision in masses, for two stars HD123999 and HD210027 we have obtained more accurate mass determinations than for any of the stars from \\cite{Torres:09::}. The accuracies of the masses of these two binaries are in the precision range covered by only close double neutron star systems characterized with radio pulsar timing. In particular, the masses for HD210027 rival in precision the mass determination of the components of the relativistic double pulsar system PSR~J0737-3039 \\citep{Nice:08::}. If our targets were all eclipsing and the accuracy in masses was limited by our RVs alone, the accuracy would be in the range 0.02\\% to 0.42\\% (the fractional accuracy of M $\\sin^3i$). The lower limit of this range is equal to the mass accuracy of PSR~B1913+16 which has the most accurate mass determination for a body outside the Solar System \\citep{Nice:08::}. In fact if we adopt our RV precision and use the achievable from the ground precision in the orbital inclination angle for eclipsing binaries, we can expect to obtain masses with a fractional precision of $\\sim$0.001\\% (see Fig.~18). Clearly, our RV technique for double-lined and eclipsing binary stars opens an exciting opportunity for derving masses of stars (and other parameters) with an unprecedented precision. These combined with parallax measurements from e.g. the planned GAIA astrometric mission and hopefully accurate abundance determinations should produce an outstanding set of parameters to tests models of the stellar structure and evolution." }, "0910/0910.2641_arXiv.txt": { "abstract": "We present an overview on how variability can be used to constrain the location of the ionized outflow in nearby Active Galactic Nuclei using high-resolution X-ray spectroscopy. Without these constraints on the location of the outflow, the kinetic luminosity and mass loss rate can not be determined. We focus on the Seyfert 1 galaxy NGC 5548, which is arguably the best studied AGN on a timescale of 10 years. Our results show that frequent observations combined with long term monitoring, such as with the \\textit{Rossi X-ray Timing Explorer (RXTE)} satellite, are crucial to investigate the effects of these outflows on their surroundings. ", "introduction": "Active Galactic Nuclei (AGN) outflows are thought to play an important role in feedback processes. The growth of the supermassive black hole is connected to the growth of the bulge and the interstellar and intergalactic medium are thought to be enriched by these AGN outflows \\cite{DiMatteo05}. However the main problem is that we do not know the location or origin of these outflows. The kinetic luminosity and mass loss rate can not be accurately determined if the location is unknown. By studying these outflows in the X-ray regime with high-resolution grating spectrometers over multiple years, we can constrain the location of the outflow by using the intrinsic variability of the source \\cite{Netzer03, Nicastro07}. We will focus on one particular Seyfert 1 galaxy, NGC 5548, because it is the best studied AGN, with high-resolution X-ray data spanning almost 8 years in total \\cite{Detmers08}. ", "conclusions": "Long-term monitoring of AGN in combination with high-resolution X-ray spectroscopy is key to constrain the location of the warm absorber outflows. Once the location has been determined, the feedback strength of the outflow (mass outflow rate, kinetic luminosity) can not be accurately determined without knowing the geometry of the outflow. General estimates can be made, either by assuming that the mass outflow rate is less than the accretion rate or by assuming a specific geometry for the outflow, but the key to unravelling the importance of feedback in these local AGN is variability, combined with multi-wavelength observations to constrain the geometry of these outflows. Only then can a full picture of feedback in local AGN be obtained. \\begin{theacknowledgments} RGD wishes to thank Elisa Costantini for useful discussions regarding the paper. SRON is supported financially by NWO, The Netherlands Organization for Scientific Research. \\end{theacknowledgments}" }, "0910/0910.0716_arXiv.txt": { "abstract": "{The photometric, structural and kinematical properties of the centers of elliptical galaxies, harbour important information of the formation history of the galaxies. In the case of non active elliptical galaxies these properties are linked in a way that surface brightness, break radius and velocity dispersion of the core lie on a fundamental plane similar to that found for their global properties.} {We construct the Core Fundamental Plane (CFP) for a sizeable sample of low redshift radio galaxies and compare it with that of non radio ellipticals.} {To pursue this aim we combine data obtained from high resolution HST images with medium resolution optical spectroscopy to derive the photometric and kinematic properties of $\\sim$40 low redshift radio galaxies. } {We find that the CFPs of radio galaxies is indistinguishable from that defined by non radio elliptical galaxies of similar luminosity. The characteristics of the CFP of radio galaxies are also consistent (same slope) with those of the Fundamental Plane (FP) derived from the global properties of radio (and non radio) elliptical galaxies. The similarity of CFP and FP for radio and non radio ellipticals suggests that the active phase of these galaxies has minimal effects for the structure of the galaxies.} {} ", "introduction": "The properties of the centers of massive elliptical galaxies are of strategic importance for the understanding of the complex processes of galaxy formation. The centers represent the bottom of the potential well of the galaxy, host massive black holes and provide a record of the past history of the galaxies. Until a decade ago the properties of the center of galaxies were very little studied because of insufficient spatial resolution of available instrumentation. Only with the use of Hubble Space Telescope, and using future large ground based telescope such a study become feasible. In the pioneering study on the nuclear properties of nearby early type galaxies (carried out with HST images) \\cite{faber} were able to probe the inner regions of a number of nearby ellipticals. They point out that the inner luminosity profiles can be parameterized by a core (or break) radius $R_b$ and a characteristic surface brightness within the break radius $<\\mu_b>$. They also showed that these parameters ($R_b$, $<\\mu_b>$) can be combined to form a Core Fundamental Plane (CFP) which is analogous to the one found for the global properties of early type galaxies (\\cite{faber}). Assuming that the cores are in dynamical equilibrium and supported by random motions, and that the velocity anisotropy does not vary too much from galaxy to galaxy, and if the core $M/L$ is a well-behaved function of any two variables $<\\mu_b>$, $R_b$, or $\\sigma_0$, then one expects that the cores of galaxies follow a law similar to that of the Fundamental Plane (FP). The main reason to study global and core properties is that it allows one to probe the mechanism of galaxy formation. On the other end in the last decade a new ingredient in this picture is represented by the discovery of the presence of a massive black bole (BH) in the centers of virtually all galaxies (e.g. \\cite{FM}, \\cite{lauer}). However, only a small fraction of these BHs exhibit associated nuclear activity (non thermal nuclear emission, X-ray and radio emission). The BHs may play an important role in the formation and evolution of massive galaxies and are also a key component for the development of the nuclear activity. Therefore the comparison of the properties of active and inactive galaxies through the FP becomes a tool to investigate the interplay between galaxy formation and nuclear activity. In a previous work, using photometrical and dynamical data for 73 low red-shift (z$<$0.2) radio galaxies (RG), we (\\cite{bettoni}) were able to compare the FP of RG with the one defined by inactive ellipticals (\\cite{jfk96}, JFK96). We showed that the same FP holds for both radio and non radio ellipticals with radio galaxies occupying the region of the most luminous and large galaxies. Till very recently few data were available on the optical nuclear properties of radio galaxies. One of the best optical sets of data is part of the study of the B2 sample of low luminosity radio galaxies (\\cite{fanti}, \\cite{Cap00}). For $\\sim$60 radio galaxies WFPC2 HST imaging is available in the $F555W$ (approximately $V$) and $F814W$ (approximately $I$) filters. A similar set of data is also available for a sample of powerful 3C radio sources. These studies revealed the presence of new and interesting features, some of them almost exclusively associated to low luminosity FR I radio galaxies. In particular, the HST observations have shown the presence of dust in a large fraction of weak (FR I) radio galaxies which takes the form of extended nuclear disks (\\cite{jaffe}, \\cite{dekoff}, \\cite{dejuan}, \\cite{verdoes}, \\cite{hans}). Such structures have been naturally identified with the reservoir of material which will ultimately accrete into the central black hole. The symmetry axis of the nuclear disk may be a useful indicator of the rotation axis of the central black hole (see e.g. \\cite{capetti1}), although the precise relationship between these two axes remains uncertain. In this paper we aim to investigate the CFP for a sample of low redshift radio galaxies and to compare it with that for radio quiet galaxies. The plan of the paper is as follows. In Section 2 we discuss our observations and data reduction methods. In Section 3 we present the CFP for our sample of RG. The implications of those observations are then discussed in Section 4. Throughout this paper we use the Concordance Cosmological Model, with $H_0=70$ kms$^{-1}$Mpc$^{-1}$, and $\\Omega_{\\Lambda}=0.70$. \\begin{table*} \\caption{The sample of low redshift B2 radio galaxies with velocity dispersion measurements} \\label{tab_b2} \\begin{center} \\begin{tabular}{lcccccccc } \\hline name & $m_V$ & z & $\\sigma_c$ & $\\Delta$$\\sigma$ & $<\\mu_b>$ & $\\Delta$$<\\mu_b>$ & $R_b$ & $\\Delta$$r_b$ \\\\ (1) & (2) & (3) & (4) & (5) & (6) & (7) & (8) & (9) \\\\ \\hline 0648+27$^+$ & 13.82 & 0.0409 & 137.8 & 38.0 & 14.94 & 0.05 & 0.26 & 0.02 \\\\ 0755+37 & 13.80 & 0.0413 & 291.0 & 17.3 & 17.05 & 0.15 & 1.09 & 0.09 \\\\ 0908+37 & 15.67 & 0.1040 & 387.1 & 18.0 & 17.36 & 0.10 & 0.50 & 0.05 \\\\ 0915+32B & 15.20 & 0.0620 & 252.1 & 24.3 & 17.00 & 0.05 & 0.52 & 0.03 \\\\ 0924+30 & 13.32 & 0.0266 & 243.7 & 14.5 & 17.33 & 0.15 & 0.96 & 0.14 \\\\ 1003+26 & 15.47 & 0.1165 & 438.4 & 21.6 & 17.94 & 0.15 & 0.41 & 0.05 \\\\ 1113+24 & 15.10 & 0.1021 & 377.8 & 10.2 & 17.44 & 0.05 & 0.47 & 0.03 \\\\ 1204+34 & 15.35 & 0.0788 & 201.0 & 29.6 & 17.47 & 0.25 & 0.46 & 0.09 \\\\ 1322+36B & 12.84 & 0.0175 & 259.1 & 12.1 & 15.45 & 0.05 & 0.62 & 0.05 \\\\ 1339+26B & 15.10 & 0.0757 & 379.1 & 18.2 & 16.41 & 0.15 & 0.32 & 0.04 \\\\ 1347+28 & 15.66 & 0.0724 & 228.0 & 19.4 & 17.18 & 0.05 & 0.34 & 0.03 \\\\ 1357+28 & 14.81 & 0.0629 & 308.3 & 11.2 & 16.80 & 0.05 & 0.40 & 0.03 \\\\ 1422+26B & 14.32 & 0.0370 & 264.9 & 11.6 & 15.44 & 0.05 & 0.13 & 0.01 \\\\ 1430+25 & 16.65 & 0.0813 & 203.6 & 14.3 & 16.72 & 0.25 & 0.13 & 0.03 \\\\ 1447+27 & 13.88 & 0.0306 & 318.4 & 11.6 & 15.77 & 0.15 & 0.47 & 0.04 \\\\ 1450+28 & 16.31 & 0.1203 & 405.6 & 23.6 & 16.18 & 0.05 & 0.16 & 0.15 \\\\ 1450+28*$^+$ & 16.31 & 0.1265 & 333.3 & 22.1 & 16.56 & 0.05 & 0.20 & 0.01 \\\\ 1502+26 & 15.42 & 0.0540 & 402.1 & 19.9 & 16.60 & 0.15 & 0.66 & 0.07 \\\\ 1512+30 & 15.37 & 0.0931 & 330.2 & 25.6 & 16.55 & 0.05 & 0.34 & 0.03 \\\\ 1521+28 & 15.09 & 0.0825 & 323.0 & 11.8 & 18.02 & 0.15 & 0.85 & 0.06 \\\\ 1527+30 & 15.64 & 0.1143 & 448.0 & 23.0 & 16.76 & 0.10 & 0.34 & 0.03 \\\\ 1553+24 & 14.41 & 0.0426 & 269.9 & 10.9 & 16.41 & 0.05 & 0.45 & 0.03 \\\\ 1557+26 & 14.97 & 0.0442 & 310.7 & 16.1 & 15.45 & 0.05 & 0.21 & 0.01 \\\\ 1613+27 & 14.90 & 0.0647 & 246.6 & 13.5 & 16.74 & 0.15 & 0.30 & 0.03 \\\\ 1658+30 & 14.47 & 0.0351 & 315.8 & 10.2 & 15.02 & 0.15 & 0.19 & 0.02 \\\\ 1726+31 & 17.04 & 0.1670 & 294.8 & 26.8 & 15.96 & 0.15 & 0.22 & 0.03 \\\\ 1827+32 & 15.10 & 0.0659 & 308.5 & 11.4 & 16.24 & 0.05 & 0.21 & 0.01 \\\\ 2236+35 & 12.57 & 0.02759 & 297.8 & 12.5 & 17.09 & 0.05 & 1.28 & 0.11 \\\\ \\hline 0034+25 & 13.35 & 0.031849 & 295.0 & 10.0 & 16.27 & 0.05 & 0.74 & 0.04 \\\\ 0055+26 & 13.80 & 0.047400 & 231.0 & 13.0 & 16.34 & 0.05 & 0.40 & 0.02 \\\\ 0120+33 & 11.43 & 0.016458 & 315.0 & 10.0 & 15.13 & 0.05 & 0.75 & 0.07 \\\\ 1217+29 & 10.41 & 0.002165 & 238.0 & 10.0 & 13.87 & 0.05 & 0.80 & 0.06 \\\\ 1256+28 & 13.88 & 0.022879 & 203.0 & 10.0 & 16.48 & 0.15 & 0.70 & 0.07 \\\\ 1257+28 & 11.94 & 0.024097 & 278.0 & 10.0 & 17.97 & 0.10 & 2.55 & 0.01 \\\\ 1525+29 & 14.95 & 0.065155 & 250.0 & 30.0 & 16.56 & 0.15 & 0.53 & 0.05 \\\\ 1610+29 & 13.29 & 0.031852 & 331.0 & 26.0 & 16.49 & 0.05 & 0.95 & 0.06 \\\\ 3C88 & 14.24 & 0.030221 & 190.0 & 22.0 & 17.47 & 0.05 & 0.98 & 0.09 \\\\ 3C192 & 15.72 & 0.059709 & 199.0 & 23.0 & 14.34 & 0.15 & 0.09 & 0.01 \\\\ 3C388 & 14.77 & 0.091700 & 365.0 & 23.0 & 18.03 & 0.15 & 0.96 & 0.09 \\\\ \\hline \\end{tabular} \\end{center} \\tiny{Columns: (1) identification of the source, (2) apparent magnitude $m_V$ from \\cite{col75} (photographic magnitudes converted to visual magnitudes, \\cite{fan78}, \\cite{gon93}, and \\cite{gon00}, (3) redshift, (4, 5) measured velocity dispersion and error in km/sec, (8 ,9), $<\\mu_b>$ and error in mag/$arcsec^2$, (8, 9) break radius $r_b$ and error in arcsecs, from \\cite{hans1};~ *companion of the radio galaxy, + not used in the CFP see Appendix A1} \\end{table*} ", "conclusions": "We have presented the photometric, structural and kinematic properties of the centers of a sample of 38 low redshift radio galaxies galaxies and have shown that the conventional parameters characterizing the centers (the break radius, its surface brightness and the central velocity dispersion) are well represented in a plane (the Core Fundamental Plane) which is indistinguishable from that of elliptical galaxies that do not exhibit radio emission. A similar result was found from the comparison of the global properties of a sample of 72 radio galaxies and local Ellipticals using the standard Fundamental Plane description (\\cite{bettoni}). The comparison of the properties of the FP, that refer to the whole galaxy, with those concering the centers (the CFP) shows that the slopes of the two planes are very similar and suggest taht the same mechanism is responsible for the link among the involved quantities. The remarkable similarity of the properties of radio and non radio elliptical galaxies in the description of both the Fundamental Plane and the Core Fundamental Plane indicates that the active phase of the galaxy connected with the strong emission at radio frequencies has likely an inconsequential effect for the structure of the whole galaxy. Moreover these results suggest that the two type of galaxies (radio and non radio) have had a similar history of formation and evolution. \\appendix" }, "0910/0910.5305_arXiv.txt": { "abstract": "We conducted radio detection observations at 8.4~GHz for 22~radio-loud broad absorption line~(BAL) quasars, selected from the Sloan Digital Sky Survey~(SDSS) Third Data Release, by a very-long-baseline interferometry~(VLBI) technique. The VLBI instrument we used was developed by the Optically ConnecTed Array for VLBI Exploration project~(OCTAVE), which is operated as a subarray of the Japanese VLBI Network~(JVN). We aimed at selecting BAL quasars with nonthermal jets suitable for measuring their orientation angles and ages by subsequent detailed VLBI imaging studies to evaluate two controversial issues of whether BAL quasars are viewed nearly edge-on, and of whether BAL quasars are in a short-lived evolutionary phase of quasar population. We detected 20 out of 22 sources using the OCTAVE baselines, implying brightness temperatures greater than 10$^{5}$~K, which presumably come from nonthermal jets. Hence, BAL outflows and nonthermal jets can be generated simultaneously in these central engines. We also found four inverted-spectrum sources, which are interpreted as Doppler-beamed, pole-on-viewed relativistic jet sources or young radio sources: single edge-on geometry cannot describe all BAL quasars. We discuss the implications of the OCTAVE observations for investigations for the orientation and evolutionary stage of BAL quasars. ", "introduction": "Broad absorption line (BAL) quasars are a subclass of active galactic nuclei~(AGNs) with rest-frame ultra-violet spectra showing absorption troughs displaced blueward from the corresponding emission lines in the high-ionization transitions of C$_\\mathrm{IV}$, Si$_\\mathrm{IV}$, N$_\\mathrm{V}$, and O$_\\mathrm{IV}$, and occasionally in low-ionization transitions, e.g., Mg$_\\mathrm{II}$ and Al$_\\mathrm{III}$ \\citep{Weymann_etal.1991}. The absorption troughs are broader than 2000~km~s$^{-1}$, sometimes as broad as $\\sim0.1c$, and are presumably due to the intervening components of the outflow originating in the activity of central engines. The fact that the most luminous quasars showing BALs more frequently \\citep{Ganguly_etal.2007} is consistent with the strong radiation-pressure-driven outflows suggested by simulation-based studies for accretion phenomena (e.g., \\cite{Proga_etal.2000,Ohsuga2007}). The observed maximum velocity of absorption as function of luminosity has an upper envelope, which can be interpreted as the terminal velocity of radiative driven wind \\citep{Ganguly_etal.2007,Laor&Brandt2002}. The intrinsic percentage of quasars with BALs is $\\sim$20\\% (e.g., \\cite{Hewett&Foltz2003}). This percentage means that the BAL phenomenon takes one of major roles in quasars'~activities. However, the principal parameter determining the finding of BAL features is still unknown. In the most widely accepted scenario, this percentage represents the covering factor of an outflowing BAL wind, which is preferentially equatorial, and the BAL features can be observed when the accretion disk is almost edge-on to the line of sight, based on spectropolarimetric measurements \\citep{Goodrich&Miller1995,Cohen_etal.1995} and a theoretical disk wind model \\citep{Murray_etal.1995}. However, some radio observations provided a counterargument to this paradigm. \\citet{Zhou_etal.2006} and \\citet{Ghosh&Punsly2007} found several BAL quasars with rapid radio variability that indicated very high brightness temperatures, which require Doppler beaming on jets with inclinations of less than \\timeform{35D}, i.e., a nearly face-on view of the accretion disk. \\citet{Becker_etal.2000} found that about one-third of the radio-detected BAL quasars showed flat radio spectra ($\\alpha>-0.5$, $S_\\nu \\propto \\nu^{\\alpha}$), preferring pole-on jets because this geometry tends to make radio sources core-dominated by significant Doppler effect only on nuclear jets. Also in optical spectropolarimetry, \\citet{Brotherton_etal.2006} found an electric vector nearly parallel to a large-scale jet axis, implying that BAL outflow is not equatorial, in a Fanaroff-Riley Class II~(FR~II) radio galaxy as a BAL quasar. Thus, a pole-on outflow would be necessary for at least some of the known radio-emitting BAL quasars. These results support an alternative proposal that BALs are not closely related to inclinations, and may be associated with a relatively short-lived (possibly episodic) evolutionary phase with a large BAL wind-covering fraction (e.g., \\cite{Briggs_etal.1984,Gregg_etal.2000}). \\citet{Gregg_etal.2006} pointed out the rarity of FR~II/BAL quasars and their observed anticorrelation between the balnicity index and radio loudness, and suggested that these properties can be explained naturally by an evolutionary scheme. \\citet{Montenegro-Montes_etal.2008} indicated that many radio-emitting BAL quasars share several radio properties common to young radio sources like Compact Steep Spectrum~(CSS) or Gigahertz-Peaked Spectrum~(GPS) sources. Thus, `inclination angle' and `evolutionary phase' are two of the most important aspects in understanding BAL quasars. Very-long-baseline interferometry~(VLBI) instruments in radio wavelengths provide exclusive and crucial opportunities to obtain spatial information about AGNs at milli-arcsecond~(mas) scales by direct measurement, which should also be useful for investigating BAL quasars. The inclination can be resolved by determining the viewing angle of the jet axis, which is supposed to be perpendicular to the accretion disk. Using the framework of Doppler beaming effects, jet axes for many AGNs have been estimated by measurements of jet asymmetry~(advancing speed, brightness, jet length, etc.) by VLBI imaging. Age as a radio source can be also estimated by measurements of its apparent linear size and expanding speed, or the age of relativistic electrons appearing on synchrotron spectrum (e.g., \\cite{Nagai_etal.2006} and references therein). VLBI observations for BAL quasars have recently begun to try to study such phenomena. \\citet{Jiang&Wang2003} observed three BAL quasars with the European VLBI Network~(EVN) at 1.6~GHz and suggested that the jet of J1556+3517 was possibly viewed from nearly pole-on because of a flat spectrum and unresolved core, while that of J1312+2319 may be far from pole-on because of the two-sided structure, and the inclination of J0957+2356 was unclear because of an unresolved steep-spectrum compact source. \\citet{Liu_etal.2008} observed for eight BAL quasar sample, including both of LoBALs and HiBALs and both of steep- and flat-spectrum sources, with the EVN$+$Multi-Element Radio Linked Interferometer Network~(MERLIN) at 1.6~GHz. High brightness temperatures and linear polarization in their core components implied a synchrotron origin for the radio emission. No systematic difference was found in the radio morphology or polarization properties between their limited number of LoBAL/HiBAL or steep/flat-spectrum sources. \\citet{Kunert-Bajraszewska&Marecki2007} observed 1045+352 with the US Very Long Baseline Array~(VLBA) at 1.7, 5, and 8.4~GHz, and found complicated radio morphology and a projected linear size of only 2.1~kpc, which suggests it might be in an early stage, and consistent with the evolutionary scenario. These investigations were still inconclusive because the detected jet structures, spatial resolutions, and frequency coverages were inadequate to definitively determine jet properties, and also because of still a small number of objects to conclude the natures of BAL quasars. In this paper, we report our VLBI observations of 22~BAL quasars at 8.4~GHz by direct measurement at the mas resolution, corresponding to the parsec scale at the distance to these sources. Our aim is to find BAL quasars that have jets suitable for determining their orientation angles and ages from jet properties for future detailed VLBI imaging studies. In Section~\\ref{section:sample}, we describe our selection processes, and we present our observations and data reduction procedures in Section~\\ref{section:observationanddatareduction}. We present the observational results in Section~\\ref{section:result}, and discuss their implications in Section~\\ref{section:discussion}. In Section~\\ref{section:summary}, we summarize the outcome of our investigation. Throughout this paper, a flat cosmology is assumed, with $H_0=70$~km~s$^{-1}$~Mpc$^{-1}$, $\\Omega_\\mathrm{M}=0.3$, and $\\Omega_\\mathrm{\\Lambda}=0.7$ \\citep{Spergel_etal.2003}. \\begin{table*} \\centering \\rotatebox[origin=c]{90}{ \\begin{minipage}[c]{1.25\\textwidth} \\caption{Optically-selected, radio-flux-limited BAL quasar sample for OCTAVE observation.}\\label{table1} \\begin{center} \\begin{tabular}{llclccrccc} \\hline\\hline \\multicolumn{1}{c}{SDSS name} & \\multicolumn{1}{c}{FIRST name} & $r_\\mathrm{FIRST}^\\mathrm{SDSS}$ & \\multicolumn{1}{c}{$z$} & $l$ & BAL type & \\multicolumn{1}{c}{$M_i^\\mathrm{SDSS}$} & $I^\\mathrm{FIRST}_\\mathrm{1.4GHz}$ & $S^\\mathrm{FIRST}_\\mathrm{1.4GHz}$ & $\\log{R*}$ \\\\ \\multicolumn{1}{c}{} & \\multicolumn{1}{c}{} & (arcsec) & \\multicolumn{1}{c}{} & (pc mas$^{-1}$) & & \\multicolumn{1}{c}{(mag)} & (mJy beam$^{-1}$) & (mJy) & \\\\ \\multicolumn{1}{c}{(1)} & \\multicolumn{1}{c}{(2)} & (3) & \\multicolumn{1}{c}{(4)} & (5) & (6) & \\multicolumn{1}{c}{(7)} & (8) & (9) & (10) \\\\\\hline SDSS J004323.43$-$001552.4 & FIRST J004323.8$-$001548 & 7.6 & 2.798 & 7.9 & H & $-$27.94 & 103 & 115 & 2.7 \\\\ SDSS J021728.62$-$005227.2 & FIRST J021728.6$-$005226 & 0.3 & 2.463 & 8.1 & H & $-$25.98 & 212 & 218 & 3.7 \\\\ SDSS J075628.24$+$371455.6 & FIRST J075628.2$+$371455 & 0.3 & 2.514 & 8.1 & HL? & $-$26.10 & 239 & 247 & 3.7 \\\\ SDSS J080016.09$+$402955.6 & FIRST J080016.0$+$402955 & 0.5 & 2.021 & 8.4 & Hi & $-$26.80 & 190 & 200 & 3.1 \\\\ SDSS J081534.16$+$330528.9 & FIRST J081534.1$+$330529 & 0.5 & 2.426 & 8.1 & nH & $-$27.04 & 328 & 342 & 3.4 \\\\ SDSS J092824.13$+$444604.7 & FIRST J092824.1$+$444604 & 0.0 & 1.904 & 8.4 & nHi & $-$27.16 & 156 & 162 & 2.8 \\\\ SDSS J100515.98$+$480533.2 & FIRST J100515.9$+$480533 & 0.2 & 2.372 & 8.2 & Hi & $-$28.29 & 206 & 209 & 2.7 \\\\ SDSS J101329.92$+$491840.9 & FIRST J101329.9$+$491841 & 0.2 & 2.201 & 8.3 & nHi & $-$26.86 & 267 & 269 & 3.3 \\\\ SDSS J101827.85$+$053030.0 & FIRST J101827.8$+$053029 & 0.2 & 1.938 & 8.4 & Hi & $-$26.62 & 284 & 297 & 3.3 \\\\ SDSS J102027.20$+$432056.2 & FIRST J102027.2$+$432056 & 0.2 & 1.962 & 8.4 & nHi & $-$26.74 & 108 & 110 & 2.9 \\\\ SDSS J103038.38$+$085324.9 & FIRST J103038.3$+$085324 & 0.6 & 1.750 & 8.5 & nHi & $-$26.57 & 108 & 172 & 3.0 \\\\ SDSS J104257.58$+$074850.5 & FIRST J104257.5$+$074850 & 0.3 & 2.665 & 8.0 & H & $-$27.32 & 374 & 382 & 3.5 \\\\ SDSS J105726.62$+$032448.0 & FIRST J105726.6$+$032448 & 0.5 & 2.832 & 7.3 & H & $-$26.90 & 138 & 157 & 3.2 \\\\ SDSS J110344.53$+$023209.9 & FIRST J110344.5$+$023209 & 0.2 & 2.514 & 8.1 & H & $-$27.63 & 163 & 166 & 2.9 \\\\ SDSS J111914.32$+$600457.2 & FIRST J111914.3$+$600457 & 0.1 & 2.646 & 8.0 & nH & $-$28.97 & 186 & 192 & 2.5 \\\\ SDSS J115944.82$+$011206.9 & FIRST J115944.8$+$011206 & 0.2 & 2.000 & 8.4 & Hi & $-$28.40 & 267 & 268 & 2.6 \\\\ SDSS J122343.16$+$503753.4 & FIRST J122343.1$+$503753 & 0.1 & 3.488 & 7.3 & H & $-$29.21 & 222 & 229 & 2.7 \\\\ SDSS J122836.92$-$030439.2 & FIRST J122836.9$-$030438 & 0.7 & 1.801 & 8.4 & nHi & $-$26.53 & 144 & 149 & 3.0 \\\\ SDSS J140507.80$+$405657.8 & FIRST J140507.7$+$405658 & 0.3 & 1.993 & 8.4 & Hi & $-$26.31 & 206 & 214 & 3.3 \\\\ SDSS J143243.29$+$410327.9 & FIRST J143243.3$+$410328 & 0.4 & 1.970 & 8.4 & nHi & $-$27.84 & 257 & 262 & 2.8 \\\\ SDSS J151005.88$+$595853.3 & FIRST J151005.4$+$595856 & 4.3 & 1.720 & 8.5 & nHi & $-$27.05 & 182 & 307 & 3.1 \\\\ SDSS J152821.65$+$531030.4 & FIRST J152821.6$+$531030 & 0.4 & 2.822 & 7.8 & H & $-$26.37 & 172 & 183 & 3.6 \\\\\\hline \\end{tabular} \\end{center} \\begin{flushleft} {\\footnotesize Col.~(1) SDSS source name; Col.~(2) Radio counter part from FIRST; Col.~(3) Difference between SDSS--FIRST positions; Col.~(4) Redshift; Col.~(5) Linear scale corresponding to 1~mas; Col.~(6) BAL subtype, listed in \\citet{Trump_etal.2006}. The code ``n'' denotes a relatively narrow trough; ``HL'' denotes a HiBAL in which broad ($\\geq$1000~km~s$^{-1}$) low-ionization absorption is also seen; ``Hi'' denotes a HiBAL-only object, in which broad low-ionization absorption is not seen even though Mg$_\\mathrm{II}$ is within the spectral coverage; and ``H'' denotes a HiBAL in which the Mg$_\\mathrm{II}$ region is not within the spectral coverage of SDSS or has a very low signal-to-noise ratio and so whether or not the object is a LoBAL as well as a HiBAL is unknown; Col.~(7) $i$-band absolute magnitude listed in \\citet{Trump_etal.2006} and calculated using a flat cosmology of $H_0=70$~km~s$^{-1}$~Mpc$^{-1}$, $\\Omega_\\mathrm{M}=0.27$, and $\\Omega_\\mathrm{\\Lambda}=0.73$ \\citep{Spergel_etal.2003}; Col.~(8) 1.4~GHz peak intensity from FIRST data with a $\\sim\\timeform{5\"}$ resolution; Col.~(9) 1.4~GHz flux density from FIRST data; Col.~(10) Radio loudness, the ratio of radio-to-optical flux densities, $R* = f_\\mathrm{5GHz}/f_\\mathrm{2500\\AA}$ (e.g., \\cite{Stocke_etal.1992}), which were calculated from $z$, $S^\\mathrm{FIRST}_\\mathrm{1.4GHz}$, $M_i$, and the assumption of a radio spectral index of $-0.5$ and an optical spectral index of $-0.5$.} \\end{flushleft} \\end{minipage} } \\end{table*} ", "conclusions": "\\label{section:discussion} \\subsection{Coexistence of nonthermal jet and BAL outflow}\\label{section:coexistence} We detected 20 radio sources using OCTAVE baselines, which assured brightness temperatures of greater than $10^5$--$10^6$~K. Although the brightness temperature of a radio supernova at a very early stage could exceed $T_\\mathrm{B}=10^6$~K (e.g., \\cite{Bietenholz_etal.2001}), its expected flux density would be far from sufficient for fringe detection for such distant quasars in our sample (cf.~\\cite{vanDyk_etal.1993}). The brightness temperatures of stellar components of luminous starbursts are $\\lesssim10^5$~K (cf. \\cite{Lonsdale_etal.1993}), which are less than those of the detected radio counterparts of BAL quasars. Also, in terms of radio luminosity, even the most radio-luminous starbursts show up to only $\\sim10^{24}$~W~Hz$^{-1}$ (e.g., \\cite{Smith_etal.1998}), in contrast to $\\sim10^{27}$--$10^{28}$~W~Hz$^{-1}$ for our BAL quasar sample. Therefore, the OCTAVE-detected radio emissions cannot be accounted for by any stellar origin. We conclude that the OCTAVE-detected radio emissions of the BAL quasars originate in nonthermal jets from AGN activity, as in the cases of other AGN radio sources. VLBI observations that previously conducted \\citep{Jiang&Wang2003,Kunert-Bajraszewska&Marecki2007,Liu_etal.2008} also support the existence of nonthermal jets in BAL quasars. The fairly high VLBI-detection rate (20/22) is evidence that BAL outflows, which are inferred from broad troughs in UV spectra, can coexist with nonthermal jets in radio-loud BAL quasars. This indicates that the accretion disks of BAL quasars can generate radiation-pressure driven strong outflows and magnetic-driven strong jets simultaneously. It is important to investigate the relationship between the properties of BAL features and radio emissions (cf.~\\cite{Ghosh&Punsly2007}) to understand accretion phenomena in quasars. \\subsection{Jet properties of observed BAL quasars --- inverted-spectrum sources}\\label{section:jetproperties:invertedspectrum} We found four inverted-spectrum ($\\alpha > 0$) sources (Column~(9) in Table~\\ref{table3}). Since optically-thin nonthermal synchrotron emission should show steep spectrum of $\\alpha \\leqq -0.5$, the observed inverted spectra should result from a lower-frequency absorption mechanism if a non-simultaneous spectral index is not affected by flux variability. Absorbed components should dominate the total spectrum for an inverted spectrum to be observed even using the FIRST beam width. Hence, an inverted spectrum suggests three possibilities: (1)~Doppler beaming effect on jets, (2)~a young~(compact) radio source, or (3)~artificially made due to flux variability, as follows. Doppler boosting can apparently enhance only optically-thick nuclear components beyond extended jets that would have been decelerated and optically-thin: the nuclear components tend to be synchrotron self-absorbed because of high-brightness temperatures, and show inverted spectrum at lower frequencies. An adequate Doppler beaming effect requires jets that are nearly aligned with our line-of-sight; this implies a face-on viewed accretion disk, which is inconsistent with the widely accepted scheme of equatorial BAL outflows. The presence of Doppler beaming effect has already been inferred from rapid flux variation between the NRAO VLA Sky Survey (NVSS; \\cite{Condon_etal.1998}) and the FIRST in several BAL quasars \\citep{Zhou_etal.2006,Ghosh&Punsly2007}. It is important to confirm the Doppler beaming in also by VLBI imaging in a future. Young radio sources also can make their spectra nuclear-dominated, because extended jets would not yet been developed. Radio sources of $\\sim100$--1000~pc or less could show peaked spectra, resulting inverted spectra possibly at $\\sim1$~GHz; the spectral-peak frequency is roughly determined by the linear size of radio structure (e.g., \\cite{Snellen_etal.2000}). The explanations of the relation between linear size and spectral-peak frequency have been suggested using synchrotron self-absorption~\\citep{Odea&Baum1997} and free--free absorption~\\citep{Bicknell_etal.1997}. \\citet{Montenegro-Montes_etal.2008} presented many radio sources of BAL quasars showing such peaked spectra. The ages could be determined from the linear size and expanding speed of radio structures; it is important to measure the expanding speed by VLBI-imaging monitor in a future. For example, in a typical VLBI scale, an apparent expanding rate of 0.1~mas yr$^{-1}$ and a linear size of two-sided structure of 100~pc leads to an estimated age of 500~yr. This situation can be adapted to the evolutionary scenario of BAL quasars in a relatively short-lived phase (e.g., \\cite{Briggs_etal.1984,Gregg_etal.2000}, and see also \\cite{Gregg_etal.2006}). Inverted spectra artificially-made due to flux variability between the observations of the OCTAVE at 8.4~GHz and the VLA at 1.4~GHz ($\\sim10$~yr) cannot be ruled out. A fraction of BAL quasars are significantly variable in the radio band \\citep{Zhou_etal.2006,Ghosh&Punsly2007}. A flux variability of $>$10\\% in 10~yr at 8.4~GHz for a 100-mJy source at $z=2$ implies $T_\\mathrm{B} > 10^{11.3}$~K, which would be above the inverse-Compton limit \\citep{Readhead1994} and require Doppler beaming effect. A flux variability of 10\\% corresponds to a change of spectral index of only $\\sim0.05$. That means that even if the observed spectra had been artificially made, we have the same conclusion that the four inverted-spectrum sources are possibly Doppler-boosted. {\\it As a result, we conclude that the inverted-spectrum sources are interpreted as Doppler-beamed, pole-on-viewed relativistic jet sources or young radio sources such as CSSs and GPSs: single edge-on geometry cannot descibe all BAL quasars.} We also check flux variation between NVSS and FIRST at 1.4~GHz for our BAL quasar sample. Using the same manner as \\citet{Ghosh&Punsly2007}, we found two sources, FIRST J075628.2$+$371455 and FIRST J122836.9$-$030438, with significant flux variation of $4.8\\sigma$ and $3.2\\sigma$; brightness temperatures were derived to be $10^{15.3}$~K and $10^{13.2}$~K, respectively. Both of the radio sources were very compact (2~mas or less) in OCTAVE observations, and the latter showed an inverted spectral index ($\\alpha=+2.0$). These situations are consistent with the presence of highly Doppler boosting on pole-on jets. It is important to confirm them also by VLBI imaging in a future. \\subsection{Jet properties of observed BAL quasars --- steep-spectrum sources}\\label{section:jetproperties:steepspectrum} The majority of the detected BAL quasars showed steep ($\\alpha<0$) radio spectra. It was unclear whether the weaker radio flux densities were observed at 8.4~GHz because extended structures were resolved out, or were due to intrinsically steep spectra on compact components. Hence, we do not discuss further in detail for the OCTAVE results. We have already stressed the importance of VLBI imaging observations for inverted-spectrum sources, and steep-spectrum sources are also needed to be VLBI-observed because they are the majority of BAL quasars and are also compact in most cases \\citep{Becker_etal.2000}. The correlated flux densities in the OCTAVE baselines were not much smaller than the $10^{-1}$-times the VLA peak intensities in most sources, despite the fact that the beam area of the OCTAVE was $\\sim10^{-5}$-times that of VLA. This indicates that radio-emitting origins considerably concentrated in parsec-scale components, which should be investigated using VLBI. As the evolutionary scenario, BAL quasars might associate young radio sources \\citep{Briggs_etal.1984,Gregg_etal.2000,Gregg_etal.2006}, which should be optically-thin small radio lobes seen in compact steep spectrum~(CSS; \\cite{Odea1998} for a review) objects. Many CSS sources have been revealed as young ($<10^5$~yr) radio galaxies by VLBI-imaging studies (e.g., \\cite{Murgia2003,Nagai_etal.2006}). VLBI-imaging studies may provide evidence supporting the evolutionary scenario for BAL quasars, if the steep-spectrum radio sources of BAL quasars are CSS objects \\citep{Kunert-Bajraszewska&Marecki2007}. \\subsection{Implications of OCTAVE observations}\\label{section:implications} Our OCTAVE observations have many implications for the study of BAL quasars. VLBI images of only four BAL quasars have been published before the OCTAVE observations \\citep{Jiang&Wang2003,Kunert-Bajraszewska&Marecki2007}. Our OCTAVE observations have dramatically increased the number of VLBI-detected BAL quasars, and have established that this AGN subclass includes a non-trivial number of radio-loud objects that can be directly imaged at mas resolutions (corresponding to parsec scales) as well as objects in the other AGN classes. The orientation angle and the age as radio sources of BAL quasars should be determined from jet properties in subsequent detailed VLBI imaging studies. The OCTAVE strongly recommends that multi-epoch and multi-frequency (and polarimetric) VLBI observations should be performed for these sources. Multi-epoch imaging can measure the proper motion of jets or lobes to estimate the inclination or kinematic age. Multi-frequency imaging can measure the spectral-index profiles along approaching and receding jets or lobes, which can discriminate between Doppler-boosted self-absorbed jets and free--free absorbed jets obscured by thermal plasma. BAL outflows could be free--free absorber along nonthermal jets; the projected one-dimensional profile of thermal BAL outflows could be studied in mas resolutions using VLBIs at multi-frequency at by measuring opacity gradient along the jets. The free--free absorber also could occur Faraday rotation to the vector of polarization axis toward jets, which is an another powerful tool to investigate the spatial profile of thermal BAL outflows. In multi-frequency VLBI observations, such processes should offer exclusive probes to the parsec-scale profile of BAL outflows. On the basis of the results of the OCTAVE observations, we are observing some of the OCTAVE sample by multi-frequency VLBI imaging. It will be important to compare the results of inclination measurements based on the VLBIs and optical spectropolarimetry, and to investigate the relation among UV/optical spectra, pc-scale geometry, and ages. Two of the most important questions in understanding BAL quasars are: (1)~Are they viewed at nearly edge-on? (2) Are they in a short-lived evolutionary phase? Our OCTAVE observations have increased the number of VLBI-detected BAL quasars, which offers more chance to determine their orientations and ages by subsequent VLBI-imaging studies to conclude the two pictures. We detected 20 out of the radio-brightest 22~sources selected from the counterparts of SDSS BAL quasars. We concluded that nonthermal pc-scale jets and thermal BAL outflows can coexist in these radio-loud BAL quasars simultaneously. BAL quasars have become targets of VLBIs to be revealed in pc scale by direct imaging, as in the cases of other AGN classes. We also found four inverted-spectrum sources, which are interpreted as Doppler-beamed, pole-on-viewed relativistic jet sources or young radio sources: single edge-on geometry cannot describe all BAL quasars. \\bigskip We thank M.~S.~Brotherton for very useful comments that improved the presentation of the paper. We also thank K.~Asada for very useful comments. This work was partially supported by a Grant-in-Aid for Young Scientists~(B; 18740107, A.~D.), a Grant-in-Aid for Scientific Research~(C; 21540250, A.~D.) and a Grant-in-Aid for Scientific Research~(C; 20540233, K.~W.) from the Japanese Ministry of Education, Culture, Sports, Science, and Technology~(MEXT). We are grateful to all the staffs and students involved in the development and operation of the OCTAVE. The OCTAVE project has been developed as a subproject under the VLBI Exploration of Radio Astrometry~(VERA) project in the National Astronomical Observatory of Japan~(NAOJ), a branch of the National Institutes of Natural Sciences~(NINS). The optical-fiber networks for OCTAVE have been provided by the GALAXY project\\footnote{The GALAXY lines connecting Usuda 64~m and Nobeyama 45~m to correlators had been closed since April 2008.} supported by NTT Corporation \\citep{Uose_etal.2002}, the Japan Gigabit Network-2 (JGN2) project operated by the National Institute of Information and Communications Technology~(NICT), and the Science Information NETwork~3~(SINET3) operated by the National Institute of Informatics~(NII). The OCTAVE array consists of six contributed antennas: the Usuda 64~m of the Japan Aerospace Exploration Agency~(JAXA), the Kashima 34~m of the NICT, the Nobeyama 45~m of the Nobeyama Radio Observatory~(NRO) in the NAOJ, the Tsukuba 32~m of the Geographical Survey Institute~(GSI), the Yamaguchi 32~m (operated by Yamaguchi University) of the NAOJ, and the Gifu 11~m of Gifu University. Time allocation and array operation for OCTAVE are being carried out under the framework of the Japanese VLBI Network~(JVN) project, which is led by the NAOJ, Hokkaido University, Gifu University, Yamaguchi University, and Kagoshima University, in cooperation with the Institute of Space and Astronautical Science~(ISAS)/JAXA, GSI and NICT. We used the US National Aeronautics and Space Administration's (NASA's) Astrophysics Data System~(ADS) abstract service, the NASA/IPAC Extragalactic Database (NED), which is operated by the Jet Propulsion Laboratory~(JPL). In addition, we used the Astronomical Image Processing System~(AIPS) software developed at the US National Radio Astronomy Observatory~(NRAO), a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc." }, "0910/0910.5133_arXiv.txt": { "abstract": "The Galactic bulge is the central spheroid of our Galaxy, containing about one quarter of the total stellar mass of the Milky Way (M$_{\\rm bulge}=1.8 \\times 10^{10} M_\\odot$; Sofue, Honma \\& Omodaka 2009). Being older than the disk, it is the first massive component of the Galaxy to have collapsed into stars. Understanding its structure, and the properties of its stellar population, is therefore of great relevance for galaxy formation models. I will review our current knowledge of the bulge properties, with special emphasis on chemical abundances, recently measured for several hundred stars. ", "introduction": "The near infrared images from the COBE/DIRBE experiment clearly showed that our Galaxy has a boxy shaped bulge (Dwek et al. 1995). Isophote deprojection revealed a barlike nature, with axes ratios 1:0.33:0.23 and with the near side on the $1^{\\rm st}$ quadrant, at $\\sim 20^\\circ$ from the Sun-Galactic center direction. These findings were later confirmed by several authors (e.g., Babusiaux \\& Gilmore 2005; Rattenbury et al. 2007a, and references therein), hence the prolate nature of the bulge is now widely accepted. The scale length of the {\\it main} bar is $\\sim 1.5$ kpc. A smaller bar (scale $\\sim 600$ pc) seems to be also present in the inner bulge (Alard 2001, Nishiyama et al. 2005), though further studies are needed to confirm and characterize this structure. A striking feature recently discovered in the outer bulge suggests that the Galactic bulge is all but a simple prolate spheroid. Along the minor axis, at distances in excess of $\\sim 700$ pc, a double red clump is clearly visible in several independent sets of data, symmetric both at positive and negative latitudes (McWilliam \\& Zoccali 2009). The double clump disappears outside the minor axis, leaving only the brighter of the two at positive longitudes, and the fainter of the two at negative longitudes. Photometric data mapping the whole bulge area are presently available only from the 2MASS survey, which is not faint enough to reach the red clump for $|b|<3^\\circ$. The {\\it Vista Variable in the Via L\\'actea} survey (Minniti et al. 2009) will solve this problem, mapping the whole bulge to much fainter magnitudes ($K_s \\sim 20$ in the coadded images), and allowing to deproject its 3D structure by means of its RR Lyrae variable stars. ", "conclusions": "The nature of the Galactic bulge is somehow puzzling. Its stellar population is old ($10-12$ Gyr) and it has a metallicity distribution compatible with chemical enrichment models assuming a fast star formation. The abundance ratio of alpha elements over iron also supports a short star formation timescale, certainly more rapid than that of the thin disk, and possibly more rapid than that of the thick disk. The presence of a radial metallicity gradient, at least outside $\\sim 600$ pc, favors a formation scenario via dissipational collapse, rather than secular evolution of the disk. Nevertheless, the bulge has the shape of a bar (perhaps including some X-shape feature in the outer part) and a cylindrical rotation velocity, both characteristics of a {\\it pseudobulge} formed via dynamical heating of a bar, resulting from disk secular evolution. A possible solution to these conflicting results may come from the confirmation of a double component bulge, as already suggested by several studies (e.g., Soto et al. 2007, Hill et al. 2009, Babusiaux et al. 2009) and seen in several bulges of external galaxies (Peletier et al. 2007). In any case, it is now clear from several independent evidences that the Galactic bulge is a complex structure. There is a metallicity gradient in the outer region that seems not to be present in the inner region. There are indications that the MDF in Baade's Window is bimodal, with each of the two component having different kinematics. Stellar ages are predicted to be different along the minor axis, compared to the edges of the bar. Even the morphology itself does not seem to be simply that of a bar, but rather something like an X-shape. The properties of the stellar population in Baade's Window cannot be considered as representative of the whole bulge: larger area photometric and spectroscopic {\\it maps} are needed in order to understand the bulge structure and origin. The VVV survey, and its spectroscopic followups, will certainly reserve many surprises in this sense." }, "0910/0910.0974.txt": { "abstract": "Due to the interaction of physics and astrophysics we are witnessing in these years a splendid synthesis of theoretical, experimental and observational results originating from three fundamental physical processes. They were originally proposed by Dirac, by Breit and Wheeler and by Sauter, Heisenberg, Euler and Schwinger. For almost seventy years they have all three been followed by a continued effort of experimental verification on Earth-based experiments. The Dirac process, $e^+e^-\\rightarrow2\\gamma$, has been by far the most successful. It has obtained extremely accurate experimental verification and has led as well to an enormous number of new physics in possibly one of the most fruitful experimental avenues by introduction of storage rings in Frascati and followed by the largest accelerators worldwide: DESY, SLAC etc. The Breit--Wheeler process, $2\\gamma\\rightarrow e^+e^-$, although conceptually simple, being the inverse process of the Dirac one, has been by far one of the most difficult to be verified experimentally. Only recently, through the technology based on free electron X-ray laser and its numerous applications in Earth-based experiments, some first indications of its possible verification have been reached. The vacuum polarization process in strong electromagnetic field, pioneered by Sauter, Heisenberg, Euler and Schwinger, introduced the concept of critical electric field $E_c=m_e^2c^3/(e\\hbar)$. It has been searched without success for more than forty years by heavy-ion collisions in many of the leading particle accelerators worldwide. The novel situation today is that these same processes can be studied on a much more grandiose scale during the gravitational collapse leading to the formation of a black hole being observed in Gamma Ray Bursts (GRBs). This report is dedicated to the scientific race. The theoretical and experimental work developed in Earth-based laboratories is confronted with the theoretical interpretation of space-based observations of phenomena originating on cosmological scales. What has become clear in the last ten years is that all the three above mentioned processes, duly extended in the general relativistic framework, are necessary for the understanding of the physics of the gravitational collapse to a black hole. Vice versa, the natural arena where these processes can be observed in mutual interaction and on an unprecedented scale, is indeed the realm of relativistic astrophysics. We systematically analyze the conceptual developments which have followed the basic work of Dirac and Breit--Wheeler. We also recall how the seminal work of Born and Infeld inspired the work by Sauter, Heisenberg and Euler on effective Lagrangian leading to the estimate of the rate for the process of electron--positron production in a constant electric field. In addition of reviewing the intuitive semi-classical treatment of quantum mechanical tunneling for describing the process of electron--positron production, we recall the calculations in \\emph{Quantum Electro-Dynamics} of the Schwinger rate and effective Lagrangian for constant electromagnetic fields. We also review the electron--positron production in both time-alternating electromagnetic fields, studied by Brezin, Itzykson, Popov, Nikishov and Narozhny, and the corresponding processes relevant for pair production at the focus of coherent laser beams as well as electron beam--laser collision. We finally report some current developments based on the general JWKB approach which allows to compute the Schwinger rate in spatially varying and time varying electromagnetic fields. We also recall the pioneering work of Landau and Lifshitz, and Racah on the collision of charged particles as well as experimental success of AdA and ADONE in the production of electron--positron pairs. We then turn to the possible experimental verification of these phenomena. We review: (A) the experimental verification of the $e^+e^-\\rightarrow 2\\gamma$ process studied by Dirac. We also briefly recall the very successful experiments of $e^+e^-$ annihilation to hadronic channels, in addition to the Dirac electromagnetic channel; (B) ongoing Earth based experiments to detect electron--positron production in strong fields by focusing coherent laser beams and by electron beam--laser collisions; and (C) the multiyear attempts to detect electron--positron production in Coulomb fields for a large atomic number $Z>137$ in heavy ion collisions. These attempts follow the classical theoretical work of Popov and Zeldovich, and Greiner and their schools. We then turn to astrophysics. We first review the basic work on the energetics and electrodynamical properties of an electromagnetic black hole and the application of the Schwinger formula around Kerr--Newman black holes as pioneered by Damour and Ruffini. We only focus on black hole masses larger than the critical mass of neutron stars, for convenience assumed to coincide with the Rhoades and Ruffini upper limit of 3.2 $M_\\odot$. In this case the electron Compton wavelength is much smaller than the spacetime curvature and all previous results invariantly expressed can be applied following well established rules of the equivalence principle. We derive the corresponding rate of electron--positron pair production and introduce the concept of dyadosphere. We review recent progress in describing the evolution of optically thick electron--positron plasma in presence of supercritical electric field, which is relevant both in astrophysics as well as ongoing laser beam experiments. In particular we review recent progress based on the Vlasov-Boltzmann-Maxwell equations to study the feedback of the created electron--positron pairs on the original constant electric field. We evidence the existence of plasma oscillations and its interaction with photons leading to energy and number equipartition of photons, electrons and positrons. We finally review the recent progress obtained by using the Boltzmann equations to study the evolution of an electron--positron-photon plasma towards thermal equilibrium and determination of its characteristic timescales. The crucial difference introduced by the correct evaluation of the role of two and three body collisions, direct and inverse, is especially evidenced. We then present some general conclusions. The results reviewed in this report are going to be submitted to decisive tests in the forthcoming years both in physics and astrophysics. To mention only a few of the fundamental steps in testing in physics we recall the starting of experimental facilities at the National Ignition Facility at the Lawrence Livermore National Laboratory as well as corresponding French Laser the Mega Joule project. In astrophysics these results will be tested in galactic and extragalactic black holes observed in binary X-ray sources, active galactic nuclei, microquasars and in the process of gravitational collapse to a neutron star and also of two neutron stars to a black hole giving origin to GRBs. The astrophysical description of the stellar precursors and the initial physical conditions leading to a gravitational collapse process will be the subject of a forthcoming report. As of today no theoretical description has yet been found to explain either the emission of the remnant for supernova or the formation of a charged black hole for GRBs. Important current progress toward the understanding of such phenomena as well as of the electrodynamical structure of neutron stars, the supernova explosion and the theories of GRBs will be discussed in the above mentioned forthcoming report. What is important to recall at this stage is only that both the supernovae and GRBs processes are among the most energetic and transient phenomena ever observed in the Universe: a supernova can reach energy of $\\sim 10^{54}$ ergs on a time scale of a few months and GRBs can have emission of up to $\\sim 10^{54}$ ergs in a time scale as short as of a few seconds. The central role of neutron stars in the description of supernovae, as well as of black holes and the electron--positron plasma, in the description of GRBs, pioneered by one of us (RR) in 1975, are widely recognized. Only the theoretical basis to address these topics are discussed in the present report. ", "introduction": "\\label{qedvac} \\emph{Quantum Electro-Dynamics} (QED), the quantum theory of electrons, positrons, and photons, was established by by Tomonaga \\cite{1946PThPh...1...27T}, Feynman \\cite{1948RvMP...20..367F,1949PhRv...76..749F,1949PhRv...76..769F}, Schwinger \\cite{1948PhRv...74.1439S,1949PhRv...75..651S,1949PhRv...76..790S} and Dyson \\cite{1949PhRv...75..486D,1949PhRv...75.1736D} and others in the 1940's and 1950's \\cite{Schwinger1958}. For decades, both theoretical computations and experimental tests have been developed to great perfection. It is now one of the fundamental pillars of the theory of the microscopic world. Many excellent monographs have been written \\cite{1959ittq.book.....B,Bjorken1998,Bjorken1965,Feynman1998,Feynman1965,1982els..book.....B,Itzykson2006,Lee1990,1951PhRv...82..664S,1954PhRv...93..615S,1954PhRv...94.1362S,Schwinger1970,Schwinger1998,1995qtf..book.....W,Kleinert1990}% , so the concepts of the theory and the techniques of calculation are well explained. On the basis of this material, we review some aspects and properties of the QED that are relevant to the subject of the present review. QED combines a relativistic extension of quantum mechanics with a quantized electromagnetic field. The nonrelativistic system has a unique ground state, which is the state with no particle, the \\emph{vacuum state\\/}. The excited states contain a fixed number of electrons and an arbitrary number of photons. As electrons are allowed to become relativistic, their number becomes also arbitrary, and it is possible to create pairs of electrons and positrons. In the modern functional integral description, the nonrelativistic system is described by a given set of fluctuating particle orbits running forward in time. If the theory is continued to an imaginary time, in which case one speaks of a \\emph{Euclidean formulation\\/}, the nonrelativistic system corresponds to a canonical statistical ensemble of trajectories. In the relativistic system, the orbits form worldlines in four-dimensional space-time which may run in any time direction, in particular they may run backwards in time, in which case the backward parts of a line correspond to positrons. The number of lines is arbitrary and the Euclidean formulation corresponds to a grand-canonical ensemble. The most efficient way of describing such an ensemble is by a single fluctuating field \\cite{Kleinert1990}. The vacuum state contains no physical particles. It does, however, harbor zero-point oscillations of the electron and photon fields. In the worldline description, the vacuum is represented by a grand-canonical ensemble of interacting closed world lines. These are called \\emph{virtual particles\\/}. Thus the vacuum contains the full complexity of a many-body problem so that one may rightfully say that \\emph{the vacuum is the world\\/} \\cite{Streater2000}. In the Fourier decomposition of the fluctuating fields, virtual particles correspond to Fourier components, or \\emph{modes\\/}, in which the 4-vectors of energy and momentum $k^{\\mu}\\equiv(k^{0},\\mathbf{k})\\equiv({\\mathcal{E}% },\\mathbf{k})$ do not satisfy the mass-shell relation \\begin{equation} k^{2}\\equiv(k^{0})^{2}-c^{2}|\\mathbf{k}|^{2}={\\mathcal{E}}^{2}-c^{2}% |\\mathbf{k}|^{2}=m_{e}^{2}c^{4}, \\label{massshellrelation}% \\end{equation} valid for real particles. The only way to evaluate physical consequences from QED is based on the smallness of the electromagnetic interaction. It is characterized by the dimensionless fine structure constant $\\alpha$. All theoretical results derived from QED are found in the form of series expansions in powers of $\\alpha$, which are expansions around the non-interacting system. Unfortunately, all these expansions are badly divergent (see e.g. Section 4.62 in \\cite{Kleinert2008}). The number of terms contributing to the same order of $\\alpha$ grows factorially fast, i.e., faster than any exponential, leading to a zero radius of convergence. Fortunately, however, the coupling $\\alpha$ is so small that the series possess an apparent convergence up to order $1/\\alpha\\approx137$, which is much higher than will be calculable for a long time to come (see e.g. Section 4.62 in \\cite{Kleinert2008}). With this rather academic limitation, perturbation expansions are well defined. In perturbation expansions, all physical processes are expressible in terms of \\emph{Feynman diagrams\\/}. These are graphic representations of the interacting world lines of all particles. Among these lines, there are some which run to infinity. They satisfy the mass shell relation (\\ref{massshellrelation}) and describe real particles observable in the laboratory. Those which remain inside a finite space-time region are virtual. The presence of virtual particles in the perturbation expansions leads to observable effects. Some of these have been measured and calculated with great accuracy. The most famous examples are \\begin{enumerate} \\item the electrostatic polarizability of quantum fluctuations of the QED vacuum has been measured in the Lamb shift \\cite{1947PhRv...72..241L,1953PhRv...89...98T}. \\item the anomalous magnetic moment of the electron \\cite{1948PhRv...73..416S,1949PhRv...76..769F,1972PhR.....3..193L,1974PhRvD..10.4007C,1977PhLB...68..191B}. \\item the dependence of the electric charge on the distance. It is observed by measuring cross-sections of electron--positron collisions, most recently in the L3-experiments at the \\emph{Large Electron-Positron Collider\\/} (LEP) at CERN \\cite{L3CERN}. \\item the \\emph{Casimir effect\\/} caused by virtual photons, i.e., by the fluctuations of the electromagnetic field in the QED vacuum \\cite{Casimir1948,Fierz1960}. It causes an attractive force \\cite{1958Phy....24..751S,1997PhRvL..78....5L,1998PhRvL..81.4549M} between two uncharged conducting plates in the vacuum (see also \\cite{2001PhRvE..63e1101H,2000PhRvA..62e2109H,2004PhRvA..69b2117C,Xue:1987fa,Xue:1988xj,1986AcPSn..34.1084X,Zheng1993}). \\end{enumerate} There are, of course, many other discussions of the effects of virtual particles caused either by external boundary conditions or by external classical fields \\cite{1992PNAS...89.4091S,1992PNAS...8911118S,1993PNAS...90..958S,1993PNAS...90.2105S,1993PNAS...90.4505S,1993PNAS...90.7285S,1994PNAS...91.6473S,Bordag1999,Bordag1996,2004JPhA...37R.209M,Milton2001,2001PhLB..508..211X,2003PhRvD..68a3004X,2003MPLA...18.1325X,2005GReGr..37..857X}. An interesting aspect of virtual particles both theoretically and experimentally is the possibility that they can become real by the effect of external fields. In this case, real particles are excited out of the vacuum. In the previous Section~\\ref{sauter} and \\ref{hew-effecitve}, we have shown that this possibility was first pointed out in the framework of quantum mechanics by Klein, Sauter, Euler and Heisenberg \\cite{1929ZPhy...53..157K,1931ZPhy...73..547S,1931ZPhy...69..742S,1936ZPhy...98..714H} who studied the behavior of the Dirac vacuum in a strong external electric field. Afterward, Schwinger studied this process and derived the probability (\\textit{Schwinger formula}) in the field theory of Quantum Electro-Dynamics, which will be described in this chapter. If the field is sufficiently strong, the energy of the vacuum can be lowered by creating an electron--positron pair. This makes the vacuum unstable. This is the \\emph{Sauter-Euler-Heisenberg-Schwinger process\\/} for electron--positron pair production. There are many reasons for the interest in the phenomenon of pair production in a strong electric field. The most compelling one is that now both laboratory conditions and astrophysical events provide possibilities for observing this process. In the following chapters, in addition to reviewing the \\textit{Schwinger formula} and QED-effective Lagrangian in constant electromagnetic fields, we will also derive the probability of pair production in an alternating field, and discuss theoretical studies of pair production in (i) electron-beam--laser collisions and (ii) superstrong Coulomb potential. In addition, the plasma oscillations of electron--positron pairs in electric fields will be reviewed in Section \\ref{time-independent}. The rest part of this chapter, we shall use natural units $\\hbar=c=1$. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \\subsection{Basic processes in Quantum Electro-Dynamics}\\label{qedprocesse} The total Lagrangian describing the interacting system of photons, electrons, and positrons reads, see e.g. \\cite{1982els..book.....B} \\begin{equation} {\\mathcal{L}}={\\mathcal{L}}_{0}^{\\gamma}+{\\mathcal{L}}_{0}^{e^{+}e^{-}% }+{\\mathcal{L}}_{\\mathrm{int}}, \\label{qcdl}% \\end{equation} where the free Lagrangians ${\\mathcal{L}}_{0}^{e^{+}e^{-}}$ and ${\\mathcal{L}% }_{0}^{\\gamma}$ for electrons and photons are expressed in terms of quantized Dirac field $\\psi(x)$ and quantized electromagnetic field $A_{\\mu}(x)$ as follows: \\begin{align} {\\mathcal{L}}_{0}^{e^{+}e^{-}} & =~\\,\\bar{\\psi}(x)(i\\gamma^{\\mu}% \\partial_{\\mu}-m_{e})\\psi(x),\\label{L0pair}\\\\ {\\mathcal{L}}_{0}^{\\gamma}~~ & =-{\\frac{1}{4}}F_{\\mu\\nu}(x)F^{\\mu\\nu }(x)+\\mathrm{gauge\\!-\\!fixing~term}. \\label{L0gamma}% \\end{align} Here $\\gamma^{\\mu}$ are the $4\\times4$ Dirac matrices, $\\bar{\\psi}% (x)\\equiv\\psi^{\\dagger}(x)\\gamma^{0}$, and $F_{\\mu\\nu}=\\partial_{\\mu}A_{\\nu }-\\partial_{\\nu}A_{\\mu}$ denotes the electromagnetic field tensor. Minimal coupling gives rise to the interaction Lagrangian \\begin{equation} {\\mathcal{L}}_{\\mathrm{int}}=-ej^{\\mu}(x)A_{\\mu}(x),\\quad\\quad j^{\\mu}% (x)=\\bar{\\psi}(x)\\gamma^{\\mu}\\psi(x). \\label{L0int}% \\end{equation} After quantization, the photon field is expanded into plane waves as \\begin{equation} A_\\mu(x)=\\int \\frac{d^3k}{2k_0(2\\pi)^3}\\sum^3_{\\lambda=1}\\left[a^{(\\lambda)}({\\bf k} )\\epsilon_\\mu^{(\\lambda)}({\\bf k} )e^{-ikx} +a^{(\\lambda)\\dagger}({\\bf k} )\\epsilon_\\mu^{(\\lambda)*}({\\bf k} )e^{ikx}\\right], \\label{photonmodes} \\end{equation} where $\\epsilon_\\mu^{(\\lambda)}$ are polarization vectors, and $a^{(\\lambda)}$, $a^{(\\lambda)\\dagger}$ are annihilation and creation operators of photons. The quantized fermion field $\\psi(x)$ has the expansion \\begin{eqnarray} \\!\\!\\!\\!\\!\\!\\!\\!\\!\\psi(x) &=& \\int{\\frac{d^3k}{ (2\\pi)^3}} \\frac{m}{ k^0} \\sum_{\\alpha=1,2}\\Big[b_\\alpha({\\bf k},s_3) u^{(\\alpha)}({\\bf k},s_3)e^{-ikx}+d^\\dagger_\\alpha({\\bf k},s_3)v^{(\\alpha)}({\\bf k},s_3)e^{ikx}\\Big],\\nonumber\\\\ \\label{fermionmodes} \\end{eqnarray} where the four-component spinors $u^{(\\alpha)}({\\bf k},s_3)$, $v^{(\\alpha)}( {\\bf k},s_3)$ are positive and negative energy solutions of the Dirac equation with momentum ${\\bf k}$ and spin component $s_3$. The operators $b({\\bf k},s_3)$, $b^\\dagger({\\bf k},s_3)$ annihilate and create electrons, the operators $d({\\bf k},s_3)$ and $d^\\dagger({\\bf k},s_3)$ do the same for positrons \\cite{1982els..book.....B}. In the framework of QED the transition probability from an initial to a final state for a given process is represented by the imaginary part of the unitary S-matrix squared \\begin{equation} {\\mathcal P}_{f\\leftarrow i}=\\left\\vert \\left\\langle \\mathrm{{f,out}\\left\\vert Im\\,{\\mathcal S}\\right\\vert {i,in}}\\right\\rangle \\right\\vert ^{2}, \\end{equation} where \\begin{equation} \\mathrm{Im\\,{\\mathcal S}}=(2\\pi)^{4}\\delta^{4}(P_{f}-P_{i})\\left\\vert M_{fi}\\right\\vert , \\end{equation} $M_{fi}$ is called matrix element and $\\delta$-function stays for energy-momentum conservation in the process. When initial state contains two particles with energies $\\epsilon_{1}$ and $\\epsilon_{2}$, and final state contain arbitrary number of particles having 3-momenta $\\mathbf{p}_{i}^{\\prime}$, the transition probability per unit time and unit volume is given by% \\begin{equation} \\frac{d{\\mathcal P}_{f\\leftarrow i}}{dVdt}=(2\\pi)^{4}\\delta^{4}(P_{f}-P_{i})\\left\\vert M_{fi}\\right\\vert ^{2}\\frac{1}{4\\epsilon_{1}\\epsilon_{2}}% %TCIMACRO{\\dprod \\limits_{i}}% %BeginExpansion {\\displaystyle\\prod\\limits_{i}} %EndExpansion \\frac{d^{3}p_{i}^{\\prime}}{\\left( 2\\pi\\right) ^{3}2\\epsilon_{i}}. \\label{dif_prob}% \\end{equation} The Lorentz invariant differential cross-section for a given process is then obtained from (\\ref{dif_prob}) by dividing it on the flux density of initial particles% \\begin{equation} d\\sigma=(2\\pi)^{4}\\delta^{4}(P_{f}-P_{i})\\left\\vert M_{fi}\\right\\vert ^{2}\\frac{1}{4I_{kin}}% %TCIMACRO{\\dprod \\limits_{i}}% %BeginExpansionv {\\displaystyle\\prod\\limits_{i}} %EndExpansion \\frac{d^{3}p_{i}^{\\prime}}{\\left( 2\\pi\\right) ^{3}2\\epsilon_{i}}, \\end{equation} where $p_{1}$ and $p_{2}$\\ are particles' 4-momenta, $m_{1}$ and $m_{2}$\\ are their masses respectively, $I_{kin}=\\sqrt{\\left( p_{1}p_{2}\\right) ^{2}-m_{1}% ^{2}m_{2}^{2}}$. It is useful to work with Mandelstam variables which are kinematic invariants built from particles 4-momenta. Consider the process $A+B\\longrightarrow C+D$. Lorentz invariant variables can be constructed in the following way% \\begin{align} s & =\\left( p_{A}+p_{B}\\right) ^{2}=\\left( p_{C}+p_{D}\\right) ^{2},\\nonumber\\\\ t & =\\left( p_{A}+p_{C}\\right) ^{2}=\\left( p_{B}+p_{D}\\right) ^{2},\\\\ u & =\\left( p_{B}+p_{C}\\right) ^{2}=\\left( p_{A}+p_{D}\\right) ^{2}.\\nonumber \\end{align} Since any incoming particle can be regarded as outgoing antiparticle, it gives rise to the crossing symmetry property of the scattering amplitude, which is best reflected in the Mandelstam variables. In fact, reactions $A+B\\longrightarrow C+D$, $A+\\bar{C}\\longrightarrow\\bar{B}+D$ or $A+\\bar {D}\\longrightarrow C+\\bar{B}$ where the bar denotes the antiparticle are just different cross-channels of a single general reaction. The meaning of the variables $s,t,u$ changes, but the amplitude is the same. The S-matrix is computed through the interaction operator as% \\begin{equation} {\\mathcal S}={\\mathcal T}\\exp\\left( i\\int{\\mathcal{L}}_{\\mathrm{int}}d^{4}x\\right) , \\end{equation} where ${\\mathcal T}$ is the chronological operator. Perturbation theory is applied, since the fine structure constant is small, while any additional interaction in collision of particles contains the factor $\\alpha$. A\\ simple and elegant way of computation of the S-matrix and consequently of the matrix element $M_{fi}$\\ is due to Feynman, who discovered a graphical way to depict each QED process, in momentum representation. In what follows we consider briefly the calculation for the case of Compton scattering process \\cite{1982els..book.....B}, which is given by two Feynman diagrams. Conservation law for 4-momenta is $p+k=p^{\\prime}+k^{\\prime}$, where $p$ and $k$ are 4-momenta of electron and photon respectively, and invariant $I_{kin}=\\frac{1}{4} (s-m_e^{2})^{2}$. After the calculation of traces with gamma-matrices, the final result, expressed in Mandelstam variables, is \\begin{align} |M_{fi}|^{2} & =2^{7}\\pi^{2}e^{4}\\left[ \\frac{m_e^{2}}{s-m_e^{2}}+\\frac{m_e^{2}% }{u-m_e^{2}}+\\left( \\frac{m_e^{2}}{s-m_e^{2}}+\\frac{m_e^{2}}{u-m_e^{2}}\\right) ^{2}\\right. \\nonumber\\\\ & \\left. -\\frac{1}{4}\\left( \\frac{s-m_e^{2}}{u-m_e^{2}}+\\frac{u-m_e^{2}}{s-m_e^{2}% }\\right) \\right] , \\label{M_fi_gamma1_1}% \\end{align} $s=(p+k)^{2}$, $t=(p-p^{\\prime})^{2}$ and$\\ u=(p-k^{\\prime})^{2}$. Since the differential cross-section is independent of the azimuth of $\\mathbf{p}% _{1}^{\\prime}$ relative to $\\mathbf{p}_{1}$,\\ it is obtained from (\\ref{M_fi_gamma1_1}) as% \\begin{equation} d\\sigma=\\frac{1}{64\\pi}|M_{fi}|^{2}\\frac{dt}{I_{kin}^{2}}. \\label{dsigma} \\end{equation} In the laboratory frame, where $s-m_e^{2}=2m_ew$, $u-m_e^{2}=-2m_ew^{\\prime}$ and electron is at rest before the collision with photon, the differential cross-section of Compton scattering is thus given by the Klein--Nishina formula \\cite{1929ZPhy...52..853K} \\begin{equation} d\\sigma=\\frac{1}{2}\\left( \\frac{e^{2}}{m}\\right) ^{2}\\left( \\frac {\\omega^{\\prime}}{\\omega}\\right) ^{2}\\left( \\frac{\\omega}{\\omega^{\\prime}}+\\frac{\\omega^{\\prime}}% {\\omega}-\\sin^{2}\\vartheta\\right) , \\label{KN}% \\end{equation} where $\\omega$ and $\\omega^{\\prime}$\\ are frequencies of photon before and after the collision, $\\vartheta$\\ is the angle at which the photon is scattered. \\subsection{The Dirac and the Breit--Wheeler processes in QED}\\label{DiracBWqed} We turn now to the formulas obtained within framework of quantum mechanics by Dirac \\cite{1930PCPS...26..361D} and Breit and Wheeler \\cite{1934PhRv...46.1087B} within QED. The crossing symmetry allows to readily write the matrix element for the pair production (\\ref{ee2gamma}) and pair annihilation (\\ref{2gammaee}) processes with the energy-momentum conservation written as $p_++p_-=k_1+k_2$, where $p_+$ and $p_-$ are 4-momenta of the positron and the electron, $k_1$ and $k_2$ are 4-momenta of two photons. It is in fact given by the same formula (\\ref{M_fi_gamma1_1}) with the substitution $p\\rightarrow p_{-},$ $p^{\\prime }\\rightarrow p_{+},$ $k\\rightarrow k_{1},$ $k^{\\prime}\\rightarrow k_{2}$, but with different meaning of the kinematic invariants $s=(p_{-}-k_{1})^{2}$, $t=(p_{-}+p_{+})^{2}$,$\\ u=(p_{-}-k_{2})^{2}$. Matrix elements for Dirac and Breit--Wheeler processes are the same. The differential cross-section of the Dirac process is obtained from (\\ref{M_fi_gamma1_1})\\ with the exchange $s\\leftrightarrow t$ and the invariant $I_{kin}=\\frac{1}{4}t(t-4m_e^{2})$, which leads to (\\ref{Dirac cross-section}). For the case of the Breit--Wheeler process with the invariant $I_{kin}=\\frac{1}{4}t^{2}$, the result is reduced to (\\ref{BW section0}). Since the Dirac pair annihilation process (\\ref{ee2gamma}) is the inverse of Breit--Wheeler pair production (\\ref{2gammaee}), it is useful to compare the cross-section of the two processes. We note that the squared transition amplitude $|M_{fi}|^2$ must be the same for two processes, due to the CPT invariance. The cross-sections could be different only by kinematics and statistical factors. Let us consider the pair annihilation process in the center of mass system where ${\\mathcal E}={\\mathcal E}_1+{\\mathcal E}_2={\\mathcal E}'_1+{\\mathcal E}'_2$ is the total energy, the initial and final momenta are equal and opposite, ${\\bf p}_1 =-{\\bf p}_2\\equiv {\\bf p}$ and ${\\bf p}'_1 =-{\\bf p}'_2\\equiv {\\bf p}'$. The differential cross-section is given by (\\ref{dsigma}). For the Breit and Wheeler process (\\ref{2gammaee}) of two colliding photons with 4-momenta $k_1$ and $k_2$, the scalar $I^2_{\\gamma\\gamma}=(k_1k_2)^2$. For the Dirac process (\\ref{ee2gamma}) of colliding electron and positron with 4-momenta $p_1$ and $p_2$, the scalar $I^2_{e^+ e^-} = (p_1p_2)^2-m_e^4$. As results, one has \\begin{equation}\\label{I} \\frac{d\\sigma_{\\gamma\\gamma}}{d\\sigma_{e^+ e^-}} =\\frac{I^2_{e^+ e^-}}{I^2_{\\gamma\\gamma}} = \\frac{2(k_1k_2)-4m_e^2}{2(k_1k_2)}= \\frac{{\\mathcal E}^2-2m_e^2}{{\\mathcal E}^2} =\\left(\\frac{|{\\bf p}|}{{\\mathcal E}}\\right)^2=\\hat\\beta^2, \\end{equation} where momenta and energies are related by \\begin{equation} (p_1+p_2)^2=(k_1+k_2)^2=2(k_1k_2)=2{\\mathcal E}^2. \\nonumber \\end{equation} Integrating Eq.~(\\ref{I}) over all scattering angles yields the total cross-section. Whereas the previous $\\sigma_{\\gamma\\gamma}$ required division by a Bose factor 2 for the two identical photons in the final state, the cross-section $\\sigma_{e^+ e^-}$ has no such factor since the final electron and positron are not identical. Hence we obtain \\begin{equation}\\label{sigmaee-sigmagg} \\sigma_{e^+ e^-} = \\frac{1}{2\\hat\\beta^2}\\sigma_{\\gamma\\gamma}. \\end{equation} By re-expressing the kinematic quantities in the laboratory frame, one obtains the Dirac cross-section (\\ref{Dirac cross-section}). As shown in Eq.~(\\ref{sigmaee-sigmagg}) in the center of mass of the system, the two cross-sections $\\sigma_{e^+ e^-}$ and $\\sigma_{\\gamma\\gamma}$ of the above described phenomena differ only in the kinematics and statistical factor $1/(2\\hat\\beta^{2})$, which is related to the fact that the resulting particles are massless or massive. The process of electron and positron production by the collision of two photons has a kinematic energy threshold, while the process of electron and positron annihilation to two photons has not such kinematic energy threshold. In the limit of high energy neglecting the masses of the electron and positron, $\\hat\\beta\\rightarrow 1$, the difference between two cross-sections $\\sigma_{e^+ e^-}$ and $\\sigma_{\\gamma\\gamma}$ is only the statistical factor $1/2$. The total cross-sections (\\ref{Dirac section0}) of Breit--Wheeler's and Dirac's process are of the same order of magnitude $\\sim 10^{-25}{\\rm cm}^2$ and have the same energy dependence $1/{\\mathcal E}^2$ above the energy threshold. The energy threshold ($2m_ec^2$) have made until now technically impossible to observe the pair production by the Breit--Wheeler process in laboratory experiments at the intersection of two beams of X-rays. Another reason is of course the smallness of the total cross-section (\\ref{bwsection3}) ($\\sigma_{\\gamma\\gamma}\\lesssim 10^{-25}$cm$^2$) and the experimental limitations on the intensities $I_i$ (\\ref{bwaa}) of the light beams. We shall see however, that this Breit--Wheeler process occurs routinely in the dyadosphere of a black hole. The observations of such phenomena in the astrophysical setting are likely to give the first direct observational test of the validity of the Breit--Wheeler process, see e.g. \\cite{Ruffini2009}. \\subsection{Double-pair production}\\label{doub} Following the Breit--Wheeler pioneer work on the process (\\ref{2gammaee}), Cheng and Wu \\cite{1969PhRvL..22..666C,1969PhRvL..23.1311C,1969PhRv..182.1852C,1969PhRv..182.1868C,1969PhRv..182.1873C,1969PhRv..182.1899C} considered the high-energy behavior of scattering amplitudes and cross-section of two photon collision, up to higher order ${\\mathcal O}(\\alpha^4)$ \\cite{1970PhRvD...1.3414C}, \\begin{equation}\\label{gamma4ee} \\gamma_{1}+\\gamma_{2}\\rightarrow e^{+}+e^{-}+ e^{+}+e^{-}. \\end{equation} For this purpose, they calculated the two photon forward scattering amplitude $M_{\\gamma\\gamma}$ (see Eq.~(\\ref{dif_prob})) by taking into account all relevant Feynman diagrams via two electron loops up to the order ${\\mathcal O}(\\alpha^4)$. The total cross-section $\\sigma_{\\gamma\\gamma}$ for photon-photon scattering is related to the photon-photon scattering amplitude $M_{\\gamma\\gamma}$ in the forward direction by the optical theorem, \\begin{equation}\\label{optict} \\sigma_{\\gamma\\gamma}(s)=\\frac{1}{s}{\\rm Im}M_{\\gamma\\gamma}, \\end{equation} where $s$ is the square of the total energy in the center of mass system. They obtained the total cross-section of double pair production (\\ref{gamma4ee}) at high energy $s\\gg 2m_e$, \\begin{equation}\\label{cwresult} \\lim_{s\\rightarrow \\infty}\\sigma_{\\gamma\\gamma}(s)= \\frac{\\alpha^4}{36\\pi m_e^2}[175\\zeta(3)-38]\\sim 6.5 \\mu b, \\end{equation} which is independent of $s$ as well as helicities of the incoming photons. Up to the $\\alpha^4$ order, Eq.~(\\ref{cwresult}) is the largest term in the total cross-section for photon-photon scattering at very high energy. This can be seen by comparing Eq.~(\\ref{cwresult}) with the cross-section (\\ref{BW section0},\\ref{bwsection3}) of the Breit--Wheeler process (\\ref{2gammaee}), which is the lowest-order process in a photon-photon collision and vanishes as $s\\rightarrow \\infty$. Thus, although the Breit--Wheeler cross-section is of lower order in $\\alpha$, the Cheng-Wu cross-section (\\ref{cwresult}) is larger as the energy becomes sufficiently high. In particular, the Cheng-Wu cross-section (\\ref{cwresult}) exceeds the Breit--Wheeler one (\\ref{BW section0}) as the center of mass energy of the photon ${\\mathcal E}\\ge 0.24$GeV. Note that in the double pair production (\\ref{gamma4ee}), the energy threshold is ${\\mathcal E}\\ge 2m_e$ rather than ${\\mathcal E}\\ge m_e$ in the one pair production of Breit--Wheeler. In Ref.~\\cite{1970PhRvD...1..467C,1970PhRvD...2.2103C}, using the same method Cheng and Wu further calculated other cross-sections for high-energy photon-photon scattering to double pion, muon and electron--positron pairs: \\begin{itemize} \\item the process of double muon pair production $\\gamma_{1}+\\gamma_{2}\\rightarrow \\mu^{+}+\\mu^{-}+ \\mu^{+}+\\mu^{-}$ and its cross-section can be obtained by replacing $m_e\\rightarrow m_\\mu$ in Eq.~(\\ref{cwresult}), thus $\\sigma_{\\gamma\\gamma}(s)\\sim 1.5\\cdot 10^{-4} \\mu b$. \\item the process of double pion pair production $\\gamma_{1}+\\gamma_{2}\\rightarrow \\pi^{+}+\\pi^{-}+ \\pi^{+}+\\pi^{-}$ and its extremely small cross-section \\begin{equation}\\label{cwpi} \\lim_{s\\rightarrow \\infty}\\sigma_{\\gamma\\gamma}(s)= \\frac{\\alpha^4}{144\\pi m_\\pi^2}[7\\zeta(3)+10]\\sim 0.23\\cdot 10^{-5} \\mu b, \\end{equation} \\item the process of one pion pair and one electron--positron pair production $\\gamma_{1}+\\gamma_{2}\\rightarrow e^{+}+e^{-}+ \\pi^{+}+\\pi^{-}$ and its small cross-section \\begin{equation}\\label{cwpiee} \\lim_{s\\rightarrow \\infty}\\sigma_{\\gamma\\gamma}(s)= \\frac{2\\alpha^4}{27\\pi m_\\pi^2}\\left[\\left(\\ln\\frac{m_\\pi^2}{m_e^2}\\right)^2 +\\frac{16}{3}\\ln\\frac{m_\\pi^2}{m_e^2}+\\frac{163}{18}\\right]\\sim 0.26\\cdot 10^{-3} \\mu b, \\end{equation} which is more than one hundred times larger than (\\ref{cwpi}). \\item the process of one pion pair and one muon-antimuon pair production $\\gamma_{1}+\\gamma_{2}\\rightarrow \\mu^{+}+\\mu^{-}+ \\pi^{+}+\\pi^{-}$ and its small cross-section can be obtained by replacing $m_e\\rightarrow m_\\mu$ in Eq.~(\\ref{cwpiee}) everywhere. \\end{itemize} \\subsection{Electron-nucleus bremsstrahlung and pair production by a photon in the field of a nucleus}\\label{bremss} The other two important QED processes, related by the crossing symmetry are the electron-nucleus bremsstrahlung (\\ref{eibr}) and creation of electron--positron pair by a photon in the field of a nucleus (\\ref{gi2p}). These processes were considered already in the early years of QED. They are of higher order, comparing to the Compton scattering and the Breit--Wheeler processes, respectively, and contain one more vertex connecting fermions with the photon. The nonrelativistic cross-section for the process (\\ref{eibr}) was derived by Sommerfeld \\cite{1931AnP...403..257S}. Here we remind the basic results obtained in the relativistic case by Bethe and Heitler \\cite{1934RSPSA.146...83B} and independently by Sauter \\cite{1934AnP...412..404S}. The Feynman diagram for bremsstrahlung can be imagined considering for the Compton scattering, but treating one of the photons as virtual one corresponding to an external field. We consider this process in Born approximation, and the momentum recoil of the nucleus is neglected. Integrating the differential cross-section over all directions of the photon and the outgoing electron one has, see e.g. \\cite{1982els..book.....B} \\begin{equation} \\begin{array}{lr} d\\sigma=Z^2\\alpha r_e^2\\frac{d\\omega}{\\omega}\\frac{p^\\prime}{p}\\left\\{\\frac{4}{3}-2\\epsilon\\epsilon^\\prime\\frac{p^2+p^{\\prime2}}{p^2 p^{\\prime2}}+m_e^2\\left(l\\frac{\\epsilon^\\prime}{p^3}+l^\\prime\\frac{\\epsilon}{p^{\\prime3}-\\frac{ll^\\prime}{pp^\\prime}}\\right)+\\right. \\\\ \\left.+L\\left[\\frac{8\\epsilon\\epsilon^\\prime}{3pp^\\prime}+\\frac{\\omega^2}{p^3p^{\\prime3}}\\left(\\epsilon^2\\epsilon^{\\prime2}+p^2p^{\\prime2}+m_e^2\\epsilon\\epsilon^\\prime\\right)+\\frac{m_e^2\\omega}{2pp^\\prime}\\left(l\\frac{\\epsilon\\epsilon^\\prime+p^2}{p^3}-l^\\prime\\frac{\\epsilon\\epsilon^\\prime+p^{\\prime2}}{p^{\\prime3}}\\right)\\right]\\right\\}, \\end{array} \\end{equation} where \\begin{equation} L=\\log\\frac{\\epsilon\\epsilon^\\prime+pp^\\prime-m_e^2}{\\epsilon\\epsilon^\\prime-pp^\\prime-m_e^2}, \\quad l=\\log\\frac{\\epsilon+p}{\\epsilon-p}, \\quad l^\\prime=\\log\\frac{\\epsilon^\\prime+p^\\prime}{\\epsilon^\\prime-p^\\prime}, \\end{equation} and $\\omega$ is photon energy, $p$ and $p^\\prime$ are electron momenta before and after the collision, respectively, $\\epsilon$ and $\\epsilon^\\prime$ are its initial and final energies. The averaged cross-section for the process of pair production by a photon in the field of a nucleus may be obtained by applying the transformation rules relating the processes (\\ref{gi2p}) and (\\ref{eibr}), see e.g. \\cite{1982els..book.....B}. The result is \\begin{equation} \\begin{array}{lcr} d\\sigma = Z^2\\alpha r_e^2\\frac{p_+p_-}{\\omega^3}d\\epsilon_+\\biggl\\{-\\frac{4}{3}-2\\epsilon_+\\epsilon_-\\frac{p_+^2+p_-^2}{p_+^2p_-^2}+m_e^2\\left(l_-\\frac{\\epsilon_+}{p_-^3}+l_+\\frac{\\epsilon_-}{p_+^3-\\frac{l_+l_-}{p_+p_-}}\\right)+ \\\\ +L\\biggl[\\frac{8\\epsilon_+\\epsilon_-}{3p_+p_-}+\\frac{\\omega^2}{p_+^3p_-^3}\\left(\\epsilon_+^2\\epsilon_-^2+p_+^2p_-^2-m_e^2\\epsilon_+\\epsilon_-\\right)- \\\\ -\\frac{m_e^2\\omega}{2p_+p_-}\\left(l_+\\frac{\\epsilon_+\\epsilon_- -p_+^2}{p_+^3}+l_-\\frac{\\epsilon_+\\epsilon_- -p_-^2}{p_-^3}\\right)\\biggr]\\biggr\\}, \\label{dsigmapm} \\end{array} \\end{equation} where \\begin{equation} L=\\log\\frac{\\epsilon_+\\epsilon_-+p_+p_-+m_e^2}{\\epsilon_+\\epsilon_- -p_+p_-+m_e^2}, \\quad l_\\pm=\\log\\frac{\\epsilon_\\pm+p_\\pm}{\\epsilon_\\pm-p_\\pm}, \\end{equation} and $p_\\pm$ and $\\epsilon_\\pm$ are momenta and energies of positron and electron respectively. The total cross-section for this process is given in Section \\ref{higher}, see (\\ref{sigmaZ1Z2}). In the ultrarelativistic approximation relaxing the condition $Z\\alpha\\ll1$ the pair production by the process (\\ref{gi2p}) was treated by Bethe and Maximon in \\cite{1954PhRv...93..768B,1954PhRv...93..788D}. The total cross-section (\\ref{sigmaZ1Z2}) is becoming then \\begin{equation} \\sigma=\\frac{28}{9}Z^2\\alpha r_e^2\\left(\\log\\frac{2\\omega}{m_e}-\\frac{109}{42}-f(Z\\alpha)\\right), \\label{sigmaZ1Z2r} \\end{equation} where \\begin{equation} f(Z\\alpha)=\\gamma_E+Re\\Psi(1+iZ\\alpha)=(Z\\alpha)^2 \\sum_{n=1}^\\infty\\frac{1}{n[n^2+(Z\\alpha)^2]}. \\end{equation} \\subsection{Pair production in collision of two ions}\\label{ionion} The process of the pair production by two ions (\\ref{2pe+e-}) in ultrarelativistic approximation was considered by Landau and Lifshitz \\cite{Landau1934} and Racah \\cite{Racah1937}. For the modern review of these topics see \\cite{2007PhR...453....1B}. The corresponding differential cross-section with logarithmic accuracy can be obtained from the differential cross-section (\\ref{dsigmapm}) taking its ultrarelativistic approximation $\\gamma\\gg1$ for the Lorentz factor of the relative motion of the two nuclei with charges $Z_1$ and $Z_2$ respectively, and treating the real photon line in the process (\\ref{gi2p}) as a virtual photon corresponding to the external field of the nucleus. One should then multiply the cross-section by the spectrum of these equivalent photons, see Section \\ref{higher}, and the result is \\begin{equation} d\\sigma=\\frac{8}{\\pi}r_e^2(Z_1Z_2\\alpha)^2\\frac{d\\epsilon_+d\\epsilon_-}{(\\epsilon_++\\epsilon_-)^4}\\left(\\epsilon_+^2+\\epsilon_-^2+\\frac{2}{3}\\epsilon_+\\epsilon_-\\right)\\log\\frac{\\epsilon_+\\epsilon_-}{m_e(\\epsilon_++\\epsilon_-)}\\log\\frac{m_e\\gamma}{(\\epsilon_++\\epsilon_-)}. \\end{equation} The total cross-section is given in Section \\ref{higher}, see (\\ref{sigmaL}) and (\\ref{sigmaR}). More recent results containing higher order in $\\alpha$ corrections are obtained in \\cite{2002PhLB..538...45B,2002PhRvA..66d2720B}, see also \\cite{2006PPNL....3..246V,2007PPNL....4...18V}. \\begin{figure}[!ht] \\centering \\includegraphics[width=5cm]{coulcor.eps} \\caption{Classification of the $e^+e^-$ pair production by the number of photons attached to a nucleus. Reproduced from \\cite{2007PhR...453....1B}.} \\label{coulcor} \\end{figure} Lepton pair production in relativistic ion collision to all orders in $Z\\alpha$ with logarithmic accuracy is studied in \\cite{2003JPhG...29.1227G} where the matrix elements are separated in different classes, see Fig. \\ref{coulcor}, according to numbers of photon lines attached to a given nucleus \\begin{equation} M=M_{(\\mathrm{i})}+M_{(\\mathrm{ii})}+M_{(\\mathrm{iii})}+M_{(\\mathrm{iv})}. \\end{equation} The Born amplitude $\\left|M_{(\\mathrm{i})}\\right|^2$ corresponding to the lowest order in $Z\\alpha$ with one photon line attached to each nucleus was computed by Landau and Lifshitz \\cite{Landau1934} who obtained the famous $L_\\gamma^3$ dependence of the cross-section in (\\ref{sigmaL}). It should be mentioned that the Racah formula (\\ref{sigmaR}) in contrast with the Landau and Lifshitz result (\\ref{sigmaL}) contains also terms proportional to $L_\\gamma^2$ and $L_\\gamma$. These terms come from the absolute square of $M_{(\\mathrm{ii})}$ and $M_{(\\mathrm{iii})}$ and their interference with the Born amplitude $M_{(\\mathrm{i})}$ \\cite{2002PhRvA..66d2720B}. Their result for the Coulomb corrections which is defined as the difference between the full cross-section and the Born approximation, in order $L_\\gamma^2$ is of the ``Bethe--Maximon'' type \\cite{2002PhRvA..66d2720B} \\begin{equation} \\sigma_\\mathrm{C}=\\frac{28}{27\\pi}r_e^2(Z_1Z_2\\alpha)^2\\left[f(Z_1\\alpha)f(Z_2\\alpha)\\right](L_\\gamma^2+{\\mathcal O}(L_\\gamma)). \\label{sigmaC} \\end{equation} The Coulomb corrections here are up to $L_\\gamma^2$-term in the Racah formula (\\ref{sigmaR}). The calculations mentioned above were all made as early as in the 1930s. Clearly, at that time only $e^+e^-$ pair production was discussed. However, these calculations can be considered for any lepton pair production, for example $\\mu^+\\mu^-$ pair production, as long as the total energy in the center of mass of the system is large enough. However, simple substitution $m_e\\longrightarrow m_l$, where $l$ stands for any lepton is not sufficient since when the energy reaches the inverse radius $\\Lambda=1/R\\sim30{\\mathrm MeV}$ of the nucleus the electric field of the nucleus cannot be approximated as a Coulomb field of a point-like particle \\cite{1998PhRvD..57.4025I}. The review of computation for lepton pair production can be found in \\cite{1973SvPhU..16..322G} and \\cite{1975PhR....15..181B}. Another effect of large enough collision energy is multiple pair production. Early work on this subject started with the observation that the impact-parameter-dependent total pair production probability computed in the lowest order perturbation theory is larger than one. The analysis of Ref. \\cite{2002PhLB..538...45B} devoted to the study of the corresponding Feynman diagrams in the high-energy limit leads to the probability of $N$ lepton pair production obeying the Poisson distribution. For a review of this topic see \\cite{2007PhR...453....1B}. Two photon particle pair production by collision of two electrons or electron and positron (see Fig. \\ref{twophoton}, where 4-momenta $p_1$ and $p_2$ correspond to their momenta) were studied in storage rings in Novosibirsk ($e^+e^-\\rightarrow e^+e^-e^+e^-$ \\cite{1971PhLB...34..663B}) and in Frascati at ADONE ($e^+e^-\\rightarrow e^+e^-e^+e^-$ \\cite{Bacci1972}, $e^+e^-\\rightarrow e^+e^-\\mu^+\\mu^-$ \\cite{1974PhRvL..32..385B}, $e^+e^-\\rightarrow e^+e^-\\pi^+\\pi^-$ \\cite{1974PhLB...48..380O}), see also \\cite{1980LNP...134....9B}. At high energy the total cross-section of two photon production of lepton pairs is given by \\cite{Landau1934,1975PhR....15..181B} \\begin{equation} \\sigma_{e^\\pm e^-\\rightarrow e^\\pm e^- l^+l^-}=\\frac{28\\alpha^4}{27\\pi m_l^2}\\left(\\ln\\frac{s}{m_e^2}\\right)^2\\ln\\frac{s}{m_e^2};\\quad l\\equiv e \\;\\; {\\mathrm or} \\;\\; \\mu. \\end{equation} see c.f. Eq. (\\ref{sigmaL}). \\subsection{QED description of pair production}\\label{qedpair} We turn now to a Sauter-Heisenberg-Euler process in QED. An external electromagnetic field is incorporated by adding to the quantum field $A_\\mu$ in (\\ref{L0int}) an unquantized external vector potential $A^{\\rm e}_\\mu$, so that the total interaction becomes \\begin{equation} {\\mathcal L}_{\\rm int}+ {\\mathcal L}^{\\rm e}_{\\rm int} = -e\\bar\\psi(x)\\gamma^\\mu \\psi(x) \\left[ A_\\mu(x)+ A^{\\rm e}_\\mu(x)\\right] . \\label{intc} \\end{equation} Instead of an operator formulation, one can derive the quantum field theory from a functional integral formulation, see e.g. \\cite{Kleinert2004}, in which the quantum mechanical partition function is described by \\begin{equation} Z[A^{\\rm e}]=\\int [{\\mathcal D}\\psi {\\mathcal D}\\bar\\psi {\\mathcal D}A_\\mu] \\exp \\left[ i\\int d^4x ({\\mathcal L} + {\\mathcal L}^{\\rm e}_{\\rm int} ) \\right], \\label{path} \\end{equation} to be integrated over all fluctuating electromagnetic and Grassmannian electron fields. The normalized quantity $Z[A^{\\rm e}]$ gives the amplitude for the vacuum to vacuum transition in the presence of the external classical electromagnetic field: \\begin{equation} \\langle {\\rm out}, 0|0, {\\rm in}\\rangle = \\frac{Z[A^{\\rm e}]}{ Z[0]}, \\label{vvamplitude} \\end{equation} where $|0, {\\rm in}\\rangle$ is the initial vacuum state at the time $t=t_-\\rightarrow -\\infty$, and $\\langle {\\rm out}, 0|$ is the final vacuum state at the time $t=t_+\\rightarrow +\\infty$. By selecting only the one-particle irreducible Feynman diagrams in the perturbation expansion of $Z[A^{\\rm e}]$ one obtains the effective action as a functional of $A^{\\rm e}$: \\begin{equation} \\Delta{\\mathcal A}_{\\rm eff}[A^{\\rm e}]\\equiv -i\\ln \\langle {\\rm out}, 0|0, {\\rm in}\\rangle. \\label{eaction} \\end{equation} In general, there exists no local effective Lagrangian density $\\Delta{\\mathcal L}_{\\rm eff}$ whose space-time integral is $\\Delta{\\mathcal A}_{\\rm eff}[A^{\\rm e}]$. An infinite set of derivatives would be needed, i.e., $\\Delta{\\mathcal L}_{\\rm eff}$ would have the arguments $A^{\\rm e}(x),\\partial_\\mu A^{\\rm e}(x), \\partial_\\mu\\partial _\\nu A^{\\rm e}(x), \\dots~$, containing gradients of arbitrary high order. With presently available methods it is possible to calculate a few terms in such a gradient expansion, or a semi-classical approximation {\\it \\`a l\\`a} JWKB for an arbitrary but smooth space-time dependence (see Section~3.21ff in Ref.~\\cite{Kleinert2004}). Under the assumption that the external field $A^{\\rm e}(x)$ varies smoothly over a finite space-time region, we may define an approximately local effective Lagrangian $\\Delta{\\mathcal L}_{\\rm eff}[A^{\\rm e}(x)]$, \\begin{equation} \\Delta{\\mathcal A}_{\\rm eff}[A^{\\rm e}]\\simeq \\int d^4x \\Delta{\\mathcal L}_{\\rm eff}[A^{\\rm e}(x)] \\approx V\\Delta t \\Delta{\\mathcal L}_{\\rm eff}[A^{\\rm e}], \\label{effl} \\end{equation} where $V$ is the spatial volume and time interval $\\Delta t=t_+-t_-$. For a large time interval $\\Delta t=t_+-t_-\\rightarrow \\infty$, the amplitude of the vacuum to vacuum transition (\\ref{vvamplitude}) has the form, \\begin{equation} \\langle {\\rm out}, 0|0 ,{\\rm in}\\rangle = e^{-i(\\Delta{\\mathcal E}_0-i\\Gamma/2)\\Delta t}, \\label{vvamplitude1} \\end{equation} where $\\Delta{\\mathcal E}_0={\\mathcal E}_0(A^{\\rm e})-{\\mathcal E}_0(0)$ is the difference between the vacuum energies in the presence and the absence of the external field, $\\Gamma$ is the vacuum decay rate, and $\\Delta t$ the time over which the field is nonzero. The probability that the vacuum remains as it is in the presence of the external classical electromagnetic field is \\begin{equation} |\\langle {\\rm out}, 0|0 ,{\\rm in}\\rangle|^2 = e^{-2{\\rm Im}\\Delta{\\mathcal A}_{\\rm eff}[A^{\\rm e}]}. \\label{probability} \\end{equation} This determines the decay rate of the vacuum in an external electromagnetic field: \\begin{equation} \\frac{ \\Gamma}{V}= \\frac{2{\\rm \\,Im}\\Delta{\\mathcal A}_{\\rm eff}[A^{\\rm e}]}{V\\Delta t} \\approx 2{\\rm \\,Im}\\Delta{\\mathcal L}_{\\rm eff}[A^{\\rm e}]. \\label{path21} \\end{equation} The vacuum decay is caused by the production of electron and positron pairs. The external field changes the energy density by \\begin{equation} \\frac{ \\Delta{\\mathcal E}_0}{V}= -\\frac{{\\rm \\,Re}\\Delta{\\mathcal A}_{\\rm eff}[A^{\\rm e}]}{V\\Delta t} \\approx -{\\rm \\,Re}\\Delta{\\mathcal L}_{\\rm eff}[A^{\\rm e}]. \\label{path21e} \\end{equation} \\subsubsection{Schwinger formula for pair production in uniform electric field}\\label{Schwingerformula} The Dirac fields appears quadratically in the partition functional (\\ref{path}) and can be integrated out, leading to \\begin{equation} Z[A^{\\rm e}]=\\int {\\mathcal D}A_\\mu\\, {\\rm Det} \\{i\\!\\!\\not{\\!\\partial}-e[\\not{\\hspace{-5pt}A(x)}+ \\not{\\hspace{-5pt}A}^{\\rm e}(x)] -m_e+i\\eta\\};\\quad \\not{\\!\\partial}\\equiv\\gamma^\\mu\\partial_\\mu, \\quad \\not{\\!A}\\equiv\\gamma^\\mu A_\\mu, \\end{equation} where Det denotes the functional determinant of the Dirac operator. Ignoring the fluctuations of the electromagnetic field, the result is a functional of the external vector potential $A^{\\rm e}(x)$: \\begin{equation} Z[A^{\\rm e}]\\approx {\\rm const}\\times {\\rm Det} \\{i\\!\\!\\not{\\!\\partial}-e \\not{\\hspace{-5pt}A}^{\\rm e}(x)-m_e+i\\eta\\}. \\label{path1} \\end{equation} The infinitesimal constant $i\\eta$ with $ \\eta >0$ specifies the treatment of singularities in energy integrals. From Eqs.~(\\ref{vvamplitude})--(\\ref{path1}), the effective action (\\ref{probability}) is given by \\begin{equation} \\Delta{\\mathcal A}_{\\rm eff}[A^{\\rm e}]=-i {\\rm Tr}\\ln \\left\\{\\left[ i\\!\\!\\not{\\!\\partial}- e\\not{\\hspace{-5pt}A}^{\\rm e}(x)-m_e+i\\eta\\right] \\frac{1}{ i\\!\\!\\not{\\!\\partial} -m_e+i\\eta}\\right\\}, \\label{eaction12} \\end{equation} where Tr denotes the functional and Dirac trace. In physical units, this is of order $\\hbar$. The result may be expressed as a one-loop Feynman diagram, so that one speaks of one-loop approximation. More convenient will be the equivalent expression \\begin{equation} \\Delta{\\mathcal A}_{\\rm eff}[A^{\\rm e}]= -\\frac{i}{2}\\,{\\rm Tr}\\ln \\left(\\{[i\\!\\!\\not{\\!\\partial}- e\\not{\\hspace{-5pt}A}^{\\rm e}(x)]^2-m_e^2+i\\eta\\} \\frac{1}{ -\\partial ^2-m_e^2+i\\eta}\\right), \\label{eaction1} \\end{equation} where \\begin{equation} [i\\!\\!\\not{\\!\\partial}- e\\not{\\hspace{-5pt}A}^{\\rm e}(x)]^2= [i\\partial_\\mu-eA^{\\rm e}_\\mu(x)]^2+\\frac{e}{2}\\sigma^{\\mu\\nu}F^{\\rm e}_{\\mu\\nu}, \\label{sqrt} \\end{equation} where $\\sigma^{\\mu\\nu}\\equiv \\frac{i}{2}[\\gamma^\\mu,\\gamma^\\nu]$,\\, $F^{\\rm e}_{\\mu\\nu}=\\partial_\\mu A^{\\rm e}_\\nu-\\partial_\\nu A^{\\rm e}_\\mu$. Using the identity \\begin{equation} \\ln\\frac{a_2}{ a_1}=\\int_0^\\infty \\frac{ds}{ s}\\big[ e^{is(a_1+i\\eta)}-e^{is(a_2+i\\eta)}\\big], \\label{identityab} \\end{equation} Eq.~(\\ref{eaction1}) becomes \\begin{equation} \\Delta{\\mathcal A}_{\\rm eff}[A^{\\rm e}] =\\frac{i}{2}\\int_0^\\infty \\frac{ds}{ s}e^{-is(m_e^2-i\\eta)}{\\rm Tr} \\langle x| e^{ is\\left\\{[i\\partial_\\mu-eA^{\\rm e}_\\mu(x)]^2+\\frac{e}{2}\\sigma^{\\mu\\nu}F^{\\rm e}_{\\mu\\nu}\\right\\}} -e^{-is\\partial ^2}|x\\rangle , \\label{probability01} \\end{equation} where $\\langle x|\\{\\cdot\\cdot\\cdot\\}|x\\rangle $ are the diagonal matrix elements in the local basis $|x\\rangle $. This is defined by the matrix elements with the momentum eigenstates $|k\\rangle$ being plane waves: $\\langle x|k\\rangle= e^{-ikx}$. The symbol ${\\rm Tr}$ denotes integral $\\int d^4x$ in space-time and the trace in spinor space. For constant electromagnetic fields, the integrand in (\\ref{probability01}) does not depend on $x$, and $\\sigma^{\\mu\\nu}F^{\\rm e}_{\\mu\\nu}$ commutes with all other operators. This will allows us to calculate the exponential in Eq.~(\\ref{probability01}) explicitly. The presence of $-i \\eta $ in the mass term ensures convergence of integral for $s\\rightarrow\\infty$. If only a constant electric field ${\\bf E}$ is present, it may be assumed to point along the ${\\hat{\\bf z}}$-axis, and one can choose a gauge such that $A^{\\rm e}_z=-Et$ is the only nonzero component of $A^{\\rm e}_\\mu$. Then one finds \\begin{equation} {\\rm tr}\\exp is\\left[\\frac{e}{2}\\sigma^{\\mu\\nu}F^{\\rm e}_{\\mu\\nu}\\right]=4\\cosh(seE), \\label{iz1} \\end{equation} where the symbol ${\\rm tr}$ denotes the trace in spinor space. Using the commutation relation $[\\partial _0,x^0]=1$, where $x^0=t$, one computes the exponential term in the effective action (\\ref{probability01}) (c.e.g.~\\cite{Itzykson2006}) \\begin{equation} \\langle x| \\exp is\\left[(i\\partial _\\mu-eA^{\\rm e}_\\mu(x))^2+\\frac{e}{2} \\sigma^{\\mu\\nu}F^{\\rm e}_{\\mu\\nu}\\right] |x\\rangle =\\frac{eE}{ (2\\pi)^2is}\\coth(eEs). \\label{iz2} \\end{equation} The second term in Eq.~(\\ref{probability01}) is obtained by setting $E= 0$ in Eq.~(\\ref{iz2}), so that the effective action (\\ref{probability01}) yields, \\begin{equation} \\Delta{\\mathcal A}_{\\rm eff}= \\frac{1}{ 2(2\\pi)^2}\\int d^4x \\int_0^\\infty \\frac{ds}{ s^3} \\left[eEs\\coth(eEs)-1\\right] e^{-is(m^2_e-i\\eta)}. \\label{oneloops1} \\end{equation} Since the field is constant, the integral over $x$ gives a volume factor, and the effective action (\\ref{probability01}) can be attributed to the space-time integral over an effective Lagrangian (\\ref{effl}) \\begin{equation} \\Delta{\\mathcal L}_{\\rm eff} = \\frac{1}{ 2(2\\pi)^2}\\int_0^\\infty \\frac{ds}{ s^3} \\left[eEs\\coth(eEs)-1\\right] e^{-is(m^2_e-i\\eta)}. \\label{oneloopl1us} \\end{equation} By expanding the integrand in Eq.~(\\ref{oneloopl1us}) in powers of $e$, one obtains, \\begin{equation} \\frac{1}{ s^3}\\left[eEs\\coth(eEs)-1\\right] e^{-i s(m^2_e-i\\eta)}= \\left[\\frac{e^2}{ 3s}E^2-\\frac{e^4s}{ 45}E^4 +{\\mathcal O}(e^6)\\right] e^{-i s(m^2_e-i\\eta)}. \\label{oneloopl1s20} \\end{equation} The small-$s$ divergence in the integrand, \\begin{equation} \\frac{e^2}{ 3}E^2 \\frac{1}{ 2(2\\pi)^2}\\int_0^\\infty \\frac{ds}{ s}e^{-is(m^2_e-i\\eta)}, \\label{oneloopl1s2} \\end{equation} is proportional to the electric term in the original Maxwell Lagrangian. The divergent term (\\ref{oneloopl1s2}) can therefore be removed by a renormalization of the field $E$. Thus we subtract a counterterm in Eq.~(\\ref{oneloopl1us}) and form \\begin{equation} \\Delta{\\mathcal L}_{\\rm eff} = \\frac{1}{ 2(2\\pi)^2}\\int_0^\\infty \\frac{ds}{ s^3} \\left[eEs\\coth(eEs)-1-\\frac{e^2}{ 3}E^2s^2\\right] e^{-is(m^2_e-i\\eta)}. \\label{oneloopl1} \\end{equation} Remembering Eq.~(\\ref{path21}), we find from (\\ref{oneloopl1}) the decay rate of the vacuum per unit volume \\begin{equation} \\frac{ \\Gamma }{V} =\\frac{1}{ (2\\pi)^2}{\\rm Im}\\int_0^\\infty \\frac{ds}{ s^3}\\left[eEs\\coth(eEs)-1-\\frac{e^2}{ 3}E^2s^2\\right] e^{-is(m_e^2-i\\eta)}. \\label{gprobability} \\end{equation} The integral (\\ref{gprobability}) can be evaluated analytically by the method of residues. Since the integrand is even, the integral can be extended to the entire $s$-axis. After this, the integration contour is deformed to enclose the negative imaginary axis and to pick up the contributions of the poles of the $\\coth$ function at $s=n\\pi/eE$. The result is \\begin{equation} \\!\\!\\!\\!\\!\\!\\!\\! \\frac{ \\Gamma }{V} =\\frac{\\alpha E^2}{ \\pi^2}\\sum_{n=1}^\\infty \\frac{1}{ n^2}\\exp \\left(-\\frac{n\\pi E_c}{ E}\\right). \\label{probability1} \\end{equation} This result, i.e. the {\\it Schwinger formula} \\cite{1951PhRv...82..664S,1954PhRv...93..615S,1954PhRv...94.1362S} is valid to lowest order in $\\hbar$ for arbitrary constant electric field strength. An analogous calculation for a charged scalar field yields \\begin{equation} \\frac{ \\Gamma }{V} =\\frac{\\alpha E^2}{ 2\\pi^2}\\sum_{n=1}^\\infty \\frac{(-1)^{n+1}}{ n^2}\\exp \\left(-\\frac{n\\pi E_c}{ E}\\right), \\label{bosonrate} \\end{equation} which generalizes the Weisskopf treatment being restricted to the leading term $n=1$. These Schwinger results complete the derivation of the probability for pair productions. The leading $n=1$ -terms of (\\ref{probability1}) and (\\ref{bosonrate}) agrees with the JWKB results we discuss in Section \\ref{semi}, and thus the correct Sauter exponential factor (\\ref{transmission}) and Heisenberg-Euler imaginary part of the effective Lagrangian (\\ref{herate}). Narozhny and Nikishov \\cite{Narozhnyi:1970uv} have expressed Eq.~(\\ref{probability}) through the probability of one pair production $P_1$, of $n$ pair production $P_n$ with $n=1,2,3,\\cdot\\cdot\\cdot$ as well as the average number of pair productions \\begin{equation} |\\langle {\\rm out}, 0|0 ,{\\rm in}\\rangle|^2 = 1- P_1-P_2-P_3-\\cdot\\cdot\\cdot , \\label{nprobability} \\end{equation} where $P_n, (n=1,2,3,\\cdot\\cdot\\cdot)$ is the probability of $n$ pair production, and the probability of one pair production is, \\begin{equation} P_1 = V\\Delta t\\frac{\\alpha E^2}{2\\pi^2}\\ln \\left(1-e^{-\\frac{\\pi m_e^2}{eE}}\\right) e^{-2V\\Delta t {\\rm Im}\\Delta{\\mathcal L}_{\\rm eff}[A^{\\rm e}]} . \\label{nprobability1} \\end{equation} The average number $\\bar N$ of pair productions is then given by \\begin{equation} \\bar N=\\sum_{n=1}^\\infty nP_n =V\\Delta t \\frac{\\alpha E^2}{ \\pi^2}\\exp \\left(-\\frac{\\pi E_c}{ E}\\right), \\label{nprobabilitym1} \\end{equation} which is the quantity directly related to the experimental measurements. \\subsubsection{Pair production in constant electromagnetic fields}\\label{EandB} Since the QED theory is gauge and Lorentz invariant, effective action $\\Delta{\\mathcal A}_{\\rm eff}$ and Lagrangian $\\Delta{\\mathcal L}_{\\rm eff}$ are expressed as functionals of the scalar and pseudoscalar invariants $S,P$ (\\ref{lightlike}). Thus they must be invariant under the discrete duality transformation: \\begin{equation} |{\\bf B}|\\rightarrow -i|{\\bf E}|,\\quad |{\\bf E}|\\rightarrow i|{\\bf B}|, \\label{duality1} \\end{equation} i.e., \\begin{equation} \\beta\\rightarrow -i\\varepsilon ,\\quad \\varepsilon\\rightarrow i\\beta. \\label{duality2} \\end{equation} This implies that in the case ${\\bf E}=0$ and ${\\bf B}\\not= 0$, results can be simply obtained by replacing $|{\\bf E}|\\rightarrow i|{\\bf B}|$ in Eqs.~(\\ref{iz2}), (\\ref{oneloopl1}), (\\ref{gprobability}): \\begin{equation} \\langle x| \\exp is\\left[(i\\partial _\\mu-eA^{\\rm e}_\\mu(x))^2+\\frac{e}{2}\\sigma^{\\mu\\nu}F^{\\rm e}_{\\mu\\nu}\\right] |x\\rangle =\\frac{eB}{ (2\\pi)^2is}\\cot(eBs), \\label{iz2b} \\end{equation} and \\begin{equation} \\Delta{\\mathcal L}_{\\rm eff} = \\frac{1}{ 2(2\\pi)^2}\\int_0^\\infty \\frac{ds}{ s^3} \\left[eBs\\cot(eBs)-1+\\frac{e^2}{ 3}B^2s^2\\right] e^{-is(m^2_e-i\\eta)}. \\label{oneloopl1b} \\end{equation} In the presence of both constant electric and magnetic fields ${\\bf E}$ and ${\\bf B}$, we adopt parallel ${\\bf E}_{\\rm CF}$ and ${\\bf B}_{\\rm CF}$ pointing along the ${\\hat{\\bf z}}$-axis in the {\\it center-of-fields frame}, as discussed after Eqs.~(\\ref{eblandau}), (\\ref{elorentz}), (\\ref{blorentz}). We can choose a gauge such that only $A^{\\rm e}_z=-E_{\\rm CF}t$, $A^{\\rm e}_y=B_{\\rm CF}x^1$ are nonzero. Due to constant fields, the exponential in the effective action Eq.~(\\ref{probability01}) can be factorized into a product of the magnetic part and the electric part. Following the same method used to compute the electric part (\\ref{iz1},\\ref{iz2}), one can compute the magnetic part by using the commutation relation $[\\partial _1,x^1]=1$, where $x^1=x$. Or one can make the substitution (\\ref{duality1}) to obtain the magnetic part, based on the discrete symmetry of duality. As results, Eqs.~(\\ref{iz1}), (\\ref{iz2}) become \\begin{equation} {\\rm tr}\\exp is\\left[\\frac{e}{2}\\sigma^{\\mu\\nu}F^{\\rm e}_{\\mu\\nu}\\right] =4\\cosh(seE_{\\rm CF})\\cos(seB_{\\rm CF}), \\label{iz1eb} \\end{equation} and \\begin{align} &\\langle x| \\exp is\\left\\{[i\\partial _\\mu-eA^{\\rm e}_\\mu(x)]^2+\\frac{e}{2}\\sigma^{\\mu\\nu}F^{\\rm e}_{\\mu\\nu}\\right\\} |x\\rangle \\nonumber\\\\ &=\\frac{1}{(2\\pi)^2}\\frac{eE_{\\rm CF}}{ is}\\coth(seE_{\\rm CF}) \\frac{eB_{\\rm CF}}{ s}\\cot(seB_{\\rm CF}). \\label{iz2eb} \\end{align} In this special frame, the effective Lagrangian is then given by \\begin{align} \\Delta{\\mathcal L}_{\\rm eff}&= \\frac{1}{ 2(2\\pi)^2}\\int_0^\\infty \\frac{ds}{ s^3} \\Big[e^2E_{\\rm CF}B_{\\rm CF} s^2\\coth(seE_{\\rm CF} ) \\cot(seB_{\\rm CF})\\nonumber\\\\ &-1-\\frac{e^2}{ 3}(E^2_{\\rm CF}-B^2_{\\rm CF})s^2 \\Big] \\cdot e^{-is(m^2_e-i\\eta)}. \\label{onelooplcf} \\end{align} Using definitions in Eqs.~(\\ref{lightlike}), (\\ref{ab}), (\\ref{fieldinvariant}), we obtain the effective Lagrangian \\begin{align} \\Delta{\\mathcal L}_{\\rm eff} &\\!=\\! \\frac{1}{ 2(2\\pi)^2}\\int_0^\\infty \\frac{ds}{ s^3} \\Big[e^2\\varepsilon\\beta s^2\\coth(e\\varepsilon s )\\cot(e\\beta s )\\nonumber\\\\ &-1-\\frac{e^2}{ 3}(\\varepsilon^2-\\beta^2)s^2\\Big] e^{-is(m^2_e-i\\eta)}; \\label{oneloopl} \\end{align} and the decay rate \\begin{align} \\frac{ \\Gamma }{V}&=\\frac{1}{ (2\\pi)^2}{\\rm Im}\\int_0^\\infty \\frac{ds}{ s^3} \\Big[e^2\\varepsilon\\beta s^2\\coth(e\\varepsilon s )\\cot(e\\beta s )\\nonumber\\\\ &-1-\\frac{e^2}{ 3}(\\varepsilon^2-\\beta^2)s^2\\Big] e^{-is(m_e^2-i\\eta)}, \\label{gprobabilityeb} \\end{align} in terms of the invariants $\\varepsilon$ and $\\beta$ (\\ref {fieldinvariant}) for arbitrary electromagnetic fields ${\\bf E}$ and ${\\bf B}$. The integral (\\ref{gprobabilityeb}) is evaluated as in Eq.~(\\ref{probability1}) by the method of residues, and yields \\cite{1951PhRv...82..664S,1954PhRv...93..615S,1954PhRv...94.1362S} \\begin{equation} \\frac{ \\Gamma }{V}=\\frac{ \\alpha \\varepsilon^2}{ \\pi^2 } \\sum_{n=1} \\frac{1}{n^2} \\frac{ n\\pi\\beta / \\varepsilon } {\\tanh {n\\pi \\beta/ \\varepsilon}}\\exp\\left(-\\frac{n\\pi E_c}{ \\varepsilon}\\right), \\label{probabilityeh} \\end{equation} which reduces for $\\beta \\rightarrow 0$ (${\\bf B}=0$) correctly to (\\ref{probability1}). The $n=1$ -term is the JWKB approximation (\\ref{wkbehfermion2}). The analogous result for bosonic fields is \\begin{equation} \\frac{ \\Gamma }{V}=\\frac{ \\alpha \\varepsilon^2}{ 2\\pi^2 } \\sum_{n=1} \\frac{(-1)^n}{n^2} \\frac{ n\\pi\\beta / \\varepsilon } {\\sinh {n\\pi \\beta/ \\varepsilon}}\\exp\\left(-\\frac{n\\pi E_c}{ \\varepsilon}\\right), \\label{probabilityehb} \\end{equation} where the first term $n=1$ corresponds to the Euler-Heisenberg result (\\ref{herate}). Note that the magnetic field produces in the fermionic case an extra factor $\\lfrac{( n\\pi\\beta / \\varepsilon )} {\\tanh ({n\\pi \\beta/ \\varepsilon}})>1$ in each term which enhances the decay rate. The bosonic series (\\ref{probabilityehb}), on the other hand, carries in each term a suppression factor $\\lfrac{ (n\\pi\\beta / \\varepsilon )} {\\sinh {n\\pi \\beta/ \\varepsilon}}<1$. The average number $\\bar N$ (\\ref{nprobabilitym1}) is given by \\begin{equation} \\bar N=\\sum_{n=1}^\\infty nP_n =V\\Delta t \\frac{\\alpha}{\\pi} \\frac{ \\alpha\\beta \\varepsilon } {\\tanh {\\pi \\beta/ \\varepsilon}}\\exp\\left(-\\frac{\\pi E_c}{ \\varepsilon}\\right). \\label{nprobabilitym} \\end{equation} The decay rate $ \\Gamma /V$ gives the number of electron--positron pairs produced per unit volume and time. The prefactor can be estimated on dimensional grounds. It has the dimension of $E_c^2/\\hbar $, i.e., $m^4 c^5/\\hbar^4$. This arises from the energy of a pair $2m_ec^2$ divided by the volume whose diameter is the Compton wavelength $\\hbar/m_ec$, produced within a Compton time $\\hbar/m_ec^2$. The exponential factor suppresses pair production as long as the electric field is much smaller than the critical electric field $E_c$, in which case the JWKB results (\\ref{wkbehfermion2}) and (\\ref{wkbehboson2}) are good approximations. The general results ({\\ref{probabilityeh}),(\\ref{probabilityehb}) was first obtained by Schwinger \\cite{1951PhRv...82..664S,1954PhRv...93..615S,1954PhRv...94.1362S} for scalar and spinor electrodynamics (see also Nikishov \\cite{Nikishov1969}, Batalin and Fradkin \\cite{1965JETP...21..375V}). The method was extended to special space-time-dependent fields in Refs.~\\cite{1971ZhPmR..13..261P,1972JETP...34..709P,2001JETPL..74..133P,Narozhnyi:1970uv,1970TMP.....5.1080B}. The monographs \\cite{Itzykson2006,Kleinert2008,Greiner1985,1980ato..book.....G,Fradkin1991} can be consulted about more detailed calculation, discussion and bibliography. \\subsubsection{Effective nonlinear Lagrangian for arbitrary constant electromagnetic field}\\label{nonlinear} Starting from the integral form of Heisenberg and Euler Lagrangian (\\ref{oneloopl}) we find explicitly real and imaginary parts of the effective Lagrangian $\\Delta{\\mathcal L}_{\\rm eff}$ (\\ref{oneloopl}) for arbitrary constant electromagnetic fields ${\\bf E}$ and ${\\bf B}$ \\cite{2006JKPSP}. The essential step is to reach a direct analytic form resulting from performing the integration. We use the expressions \\cite{1994tisp.book.....G}, \\begin{align} e\\varepsilon s\\coth(e\\varepsilon s ) &=\\sum_{n=-\\infty}^{\\infty}\\frac{s^2 }{(s^2+\\tau^2_n)};\\quad \\tau_n\\equiv n\\pi/e\\varepsilon,\\label{cosnm0}\\\\ e\\beta s\\cot(e\\beta s ) &=\\sum_{m=-\\infty}^{\\infty}\\frac{s^2 }{(s^2-\\tau^2_m)},\\quad \\tau_m\\equiv m\\pi/e\\beta, \\label{cosnm} \\end{align} and obtain for the finite effective Lagrangian of Heisenberg and Euler integral representation, \\begin{align} \\Delta{\\mathcal L}_{\\rm eff} &\\!=\\! \\frac{1}{ 2(2\\pi)^2}\\sum_{n,m=-\\infty}^{\\infty}{\\!\\!\\!}'\\int_0^\\infty ds\\frac{s}{ \\tau^2_n+\\tau^2_m} \\Big[\\frac{\\bar\\delta_{m0} }{(s^2-\\tau^2_m)}-\\frac{\\bar\\delta_{n0} }{(s^2+\\tau^2_n)}\\Big] e^{-is(m^2_e-i\\eta)}, \\label{onelooplfini} \\end{align} where divergent terms $n\\not=0,m=0$, $n=0,m\\not=0$ and $n=m=0$ are excluded from the sum, as indicated by a prime. The symbol $\\bar\\delta_{ij}\\equiv 1-\\delta_{ij}$ denotes the complimentary Kronecker-$\\delta$ which vanishes for $i=j$. The divergent term with $n=m=0$ is eliminated by the zero-field subtraction in Eq.~(\\ref{oneloopl}), while the divergent terms $n\\not=0,m=0$ and $n=0,m\\not=0$ \\begin{align} \\Delta{\\mathcal L}^{\\rm div}_{\\rm eff} = \\frac{1}{ 2(2\\pi)^2}\\int_0^\\infty \\frac{ds}{ s}e^{-is(m^2_e-i\\eta)}2\\left( \\sum_{m=1}^{\\infty}\\frac{1}{\\tau^2_m}-\\sum_{n=1}^{\\infty} \\frac{1}{\\tau^2_n}\\right), \\label{div} \\end{align} are eliminated by the second subtraction in Eq.~(\\ref{oneloopl}). This can be seen by performing the sums \\begin{equation} \\sum_{n=1}^\\infty\\frac{1}{\\tau^k_n}=\\left(\\frac{e\\varepsilon}{\\pi}\\right)^k\\zeta(k);\\quad \\sum_{n=1}^\\infty\\frac{1}{\\tau^k_m}=\\left(\\frac{e\\beta}{\\pi}\\right)^k\\zeta(k), \\label{sumrule} \\end{equation} where $\\zeta(k)=\\sum_n 1/n^k$ is the Riemann function. The infinitesimal $i\\eta$ accompanying the mass term in the $s$-integral (\\ref{onelooplfini}) is equivalent to replacing $e^{-is(m^2_e-i\\eta)}$ by $e^{-is(1-i\\eta)m^2_e}$. This implies that $s$ is to be integrated slightly below (above) the real axis for $s>0$ ($s<0$). Equivalently one may shift the $\\tau_m$ ($-\\tau_m$) variables slightly upwards (downwards) to $\\tau_m+i\\eta$ ($-\\tau_m-i\\eta$) in the complex plane. In order to calculate the finite effective Lagrangian (\\ref{onelooplfini}), the factor $e^{-is(1-i\\eta)m^2_e}$ is divided into its sin and cos parts: \\begin{align} \\Delta{\\mathcal L}^{\\sin}_{\\rm eff} &= \\frac{-i}{ 4(2\\pi)^2}\\sum_{n,m=-\\infty}^{\\infty}{\\!\\!\\!}'\\int_{-\\infty}^\\infty \\frac{sds}{ \\tau^2_n+\\tau^2_m} \\Big[\\frac{\\bar\\delta_{m0} }{(s^2-\\tau^2_m)}-\\frac{\\bar\\delta_{n0} }{(s^2+\\tau^2_n)}\\Big]\\sin[s(1-i\\eta)m^2_e]; \\label{onelooplsin}\\\\ \\Delta{\\mathcal L}^{\\cos}_{\\rm eff} &= \\frac{1}{ 2(2\\pi)^2}\\sum_{n,m=-\\infty}^{\\infty}{\\!\\!\\!}'\\int_0^\\infty \\frac{sds}{ \\tau^2_n+\\tau^2_m} \\Big[\\frac{\\bar\\delta_{m0} }{(s^2-\\tau^2_m)}-\\frac{\\bar\\delta_{n0} }{(s^2+\\tau^2_n)}\\Big] \\cos[s(1-i\\eta)m^2_e]. \\label{onelooplcos0} \\end{align} The sin part (\\ref{onelooplsin}) has an even integrand allowing for an extension of the $s$-integral over the entire $s$-axis. The contours of integration can then be closed by infinite semicircles in the half plane, the integration receives contributions from poles $\\pm \\tau_m, \\pm i\\tau_n$, so that the residue theorem leads to, \\begin{align} \\Delta{\\mathcal L}^{\\sin}_{\\rm eff} &= i\\frac{\\alpha\\varepsilon\\beta }{ 2\\pi} \\sum^\\infty_{n=1}\\frac{1}{n}\\coth\\left(\\frac{n\\pi \\beta}{ \\varepsilon}\\right) \\exp(-n\\pi E_c/\\varepsilon)\\label{sinE}\\\\ &-i\\frac{\\alpha\\varepsilon\\beta }{ 2\\pi}\\sum^\\infty_{m=1}\\frac{1}{m}\\coth\\left(\\frac{m\\pi \\varepsilon}{ \\beta}\\right) \\exp (im\\pi E_c/\\beta)\\label{sinB} \\end{align} The first part (\\ref{sinE}) leads to the exact non-perturbative Schwinger rate (\\ref{probabilityeh}) for pair production. The second term, as we see below, is canceled by the imaginary part of the cos term. In fact, shifting $s\\rightarrow s-i\\eta$, we rewrite the cos part of effective Lagrangian (\\ref{onelooplcos0}) as \\begin{equation} \\Delta{\\mathcal L}^{\\cos}_{\\rm eff} =\\frac{1}{2(2\\pi)^2} \\sum_{n,m=-\\infty}^{\\infty}{\\!\\!\\!}' \\int_0^\\infty ds \\frac{\\cos(sm^2_e)}{\\tau^2_n+\\tau^2_m} \\left(\\frac{s\\bar\\delta_{m0} }{ s^2-\\tau^2_m-i\\eta}-\\frac{s\\bar\\delta_{n0} }{ s^2+\\tau^2_n-i\\eta}\\right). \\label{onelooplcos} \\end{equation} In the first term of magnetic part, singularities $s=\\tau_m, (m>0)$ and $s=-\\tau_m, (m<0)$ appear in integrating $s$-axis. We decompose \\begin{align} \\frac{s}{ s^2-\\tau^2_m-i\\eta}=i\\frac{\\pi}{2}\\delta(s-\\tau_m)+i\\frac{\\pi}{2}\\delta(s+\\tau_m) +{\\mathcal P}\\frac{s}{ s^2-\\tau^2_m},\\label{h+} \\end{align} where ${\\mathcal P}$ indicates the principle value under the integral. The integrals over the $\\delta$-functions give \\begin{equation} \\Delta_\\delta{\\mathcal L}^{\\cos}_{\\rm eff}=i\\frac{\\alpha\\varepsilon\\beta }{ 2\\pi}\\sum^\\infty_{m=1}\\frac{1}{m}\\coth\\left(\\frac{m\\pi \\varepsilon}{ \\beta}\\right) \\exp (im\\pi E_c/\\beta), \\label{hagensin} \\end{equation} which exactly cancels the second part (\\ref{sinB}) of the sin part $\\Delta{\\mathcal L}^{\\sin}_{\\rm eff}$. It remains to find the principle-value integrals in Eq.~(\\ref{onelooplcos}), which corresponds to the real part of the effective Lagrangian \\begin{equation} (\\Delta{\\mathcal L}^{\\cos}_{\\rm eff})_{\\mathcal P} =\\frac{1}{2(2\\pi)^2} \\sum_{n,m=-\\infty}^{\\infty}{\\!\\!\\!}' \\frac{1}{\\tau^2_n+\\tau^2_m}{\\mathcal P}\\int_0^\\infty ds \\cos(sm^2_e) \\left(\\frac{s\\bar\\delta_{m0} }{ s^2-\\tau^2_m}-\\frac{s\\bar\\delta_{n0} }{ s^2+\\tau^2_n}\\right). \\label{onelooplcosp} \\end{equation} We rewrite the cos function as $\\cos(sm^2_e)=(e^{ism^2_e}+e^{-ism^2_e})/2$ and make the rotations of integration contours by $\\pm\\pi/2$ respectively, \\begin{eqnarray} (\\Delta{\\mathcal L}^{\\cos}_{\\rm eff})_{\\mathcal P} &=&\\frac{1}{2(2\\pi)^2} \\sum_{n,m=-\\infty}^{\\infty}{\\!\\!\\!}' \\frac{1}{\\tau^2_n+\\tau^2_m}\\times\\nonumber\\\\ &\\times&{\\mathcal P}\\int_0^\\infty d\\tau \\left(\\frac{\\bar\\delta_{m0}\\tau e^{-\\tau} }{ \\tau^2-(i\\tau_mm_e^2)^2} -\\frac{\\bar\\delta_{n0}\\tau e^{-\\tau} }{ \\tau^2-(\\tau_nm_e^2)^2}\\right). \\label{onelooplcosp1} \\end{eqnarray} Using the formulas (see Secs.~3.354, 8.211.1 and 8.211.2 in Ref.~\\cite{1994tisp.book.....G}) \\begin{equation} J(z) \\equiv {\\mathcal P}\\int^\\infty_0 ds \\frac{s e^{-s}}{ s^2-z^2} = -\\frac{1}{2}\\Big[e^{-z}{\\rm Ei}(z) + e^{z}{\\rm Ei}(-z)\\Big], \\label{J(z)1} \\end{equation} where ${\\rm Ei}(z)$ is the exponential-integral function, \\begin{equation} {\\rm Ei}(z) \\equiv {\\mathcal P}\\int_{-\\infty}^z dt \\frac{e^t}{t}=\\log(-z)+\\sum_{k=1}^\\infty\\frac{z^k}{kk!}, \\label{J(z)2} \\end{equation} we obtain the principal-value integrals (\\ref{onelooplcosp1}), \\begin{equation} (\\Delta{\\mathcal L}^{\\cos}_{\\rm eff})_{\\mathcal P} = \\frac{1}{2(2\\pi)^2}\\sum_{n,m=-\\infty}^{\\infty}{\\!\\!\\!}' \\frac{1}{ \\tau^2_m+\\tau^2_n} \\Big[\\bar\\delta_{m0}J(i\\tau_m m^2_e)-\\bar\\delta_{n0}J(\\tau_n m^2_e)\\Big]. \\label{pertur} \\end{equation} Having so obtained the real part of an effective Lagrangian for an arbitrary constant electromagnetic field we recover the usual approximate results by suitable expansion of the exact formula. With the help of the series and asymptotic representation (see formula 8.215 in Ref.~\\cite{1994tisp.book.....G}) of the exponential-integral function ${\\rm Ei}(z)$ for large $z$, corresponding to weak electromagnetic fields ($\\varepsilon\\ll 1, \\beta\\ll 1$), \\begin{equation} J(z) =-\\frac{1}{z^2}-\\frac{6}{z^4}-\\frac{120}{z^6}-\\frac{5040}{z^8}-\\frac{362880}{z^{10}}+\\cdot\\cdot\\cdot , \\label{J(z)3} \\end{equation} and Eq.~(\\ref{pertur}), we find, \\begin{align} (\\Delta{\\mathcal L}^{\\cos}_{\\rm eff})_{\\mathcal P} &= \\frac{1}{2(2\\pi)^2}\\sum_{n,m=-\\infty}^{\\infty}{\\!\\!\\!}' \\frac{1}{ \\tau^2_m+\\tau^2_n} \\Big\\{\\bar\\delta_{n0}\\left[\\frac{1}{\\tau_n^2m_e^4}+\\frac{6}{\\tau_n^4m_e^8}+\\frac{120}{\\tau_n^6m_e^{12}}+\\cdot\\cdot\\cdot\\right]\\nonumber\\\\ &+\\bar\\delta_{m0}\\left[\\frac{1}{\\tau_m^2m_e^4}-\\frac{6}{\\tau_m^4m_e^8}+\\frac{120}{\\tau_m^6m_e^{12}}+\\cdot\\cdot\\cdot\\right]\\Big\\}. \\label{perexpansion} \\end{align} Applying the summation formulas (\\ref{sumrule}), the weak field expansion (\\ref{perexpansion}) is seen to agree with the Heisenberg and Euler effective Lagrangian \\cite{1936ZPhy...98..714H}, \\begin{align} (\\Delta{\\mathcal L}_{\\rm eff})_{\\mathcal P} &= \\frac{2 \\alpha}{90 \\pi E_c^2}\\left\\{ ({\\bf E}^2\\!-\\!{\\bf B}^2)^2+7 ({\\bf E}\\cdot {\\bf B})^2 \\right\\}\\nonumber\\\\ &+\\frac{2 \\alpha}{315 \\pi^2 E_c^4}\\left\\{ 2({\\bf E}^2\\!-\\!{\\bf B}^2)^3+13({\\bf E}^2\\!-\\!{\\bf B}^2) ({\\bf E}\\cdot {\\bf B})^2 \\right\\}+\\cdot\\cdot\\cdot , \\label{Kleinert1} \\end{align} which is expressed in terms of a powers series of weak electromagnetic fields up to $O(\\alpha^3)$. The expansion coefficients of the terms of order $n$ have the general form $m_e^4/(E_c)^n$. As long as the fields are much smaller than $E_c$, the expansion converges. On the other hand, we can address the limiting form of the effective Lagrangian (\\ref{pertur}) corresponding to electromagnetic fields ($\\varepsilon\\gg 1, \\beta\\gg 1$). We use the series and asymptotic representation (formulas 8.214.1 and 8.214.2 in Ref.~\\cite{1994tisp.book.....G}) of the exponential-integral function ${\\rm Ei}(z)$ for small $z\\ll 1$, \\begin{equation} J(z) =-\\frac{1}{2}\\Big[e^z\\ln(z)+e^{-z}\\ln(-z)\\Big]+\\gamma_E +{\\mathcal O}(z), \\label{smallz} \\end{equation} with $\\gamma_E=0.577216$ being the Euler-Mascheroni constant, we obtain the leading terms in the strong field expansion of Eq.~(\\ref{pertur}), \\begin{equation} (\\Delta{\\mathcal L}^{\\rm cos}_{\\rm eff})_{\\mathcal P} =\\frac{1}{2(2\\pi)^2}\\sum_{n,m=-\\infty}^{\\infty '} \\frac{1}{ \\tau^2_m+\\tau^2_n}\\Big[\\bar\\delta_{n0}\\ln(\\tau_n m^2_e)-\\bar\\delta_{m0}\\ln(\\tau_m m^2_e)\\Big]+\\cdot\\cdot\\cdot . \\label{strongexp} \\end{equation} In the case of vanishing magnetic field ${\\bf B}=0$ and $m=0$ in Eq.~(\\ref{strongexp}), we have, \\begin{equation} (\\Delta{\\mathcal L}^{\\rm cos}_{\\rm eff})_{\\mathcal P} =\\frac{1}{2(2\\pi)^2}\\sum_{n=1}^{\\infty} \\frac{1}{ \\tau^2_n} \\ln(\\tau_n m^2_e)+\\cdot\\cdot\\cdot =\\frac{\\alpha E^2}{2\\pi^2}\\sum_{n=1}^{\\infty} \\frac{1}{ n^2} \\ln\\left(\\frac{n\\pi E_c}{E}\\right)+\\cdot\\cdot\\cdot , \\label{strongexpe} \\end{equation} for a strong electric field $\\bf E$. In the case of vanishing electric field ${\\bf E}=0$ and $n=0$ in Eq.~(\\ref{strongexp}), we obtain for the strong magnetic field $\\bf B$, \\begin{equation} (\\Delta{\\mathcal L}^{\\rm cos}_{\\rm eff})_{\\mathcal P} =-\\frac{1}{2(2\\pi)^2}\\sum_{m=1}^{\\infty} \\frac{1}{ \\tau^2_m} \\ln(\\tau_m m^2_e)+\\cdot\\cdot\\cdot =-\\frac{\\alpha B^2}{2\\pi^2}\\sum_{m=1}^{\\infty} \\frac{1}{ m^2} \\ln\\left(\\frac{m\\pi E_c}{B}\\right)+\\cdot\\cdot\\cdot . \\label{strongexpb} \\end{equation} The ($m=1$) term is the one obtained by Weisskopf \\cite{Weisskopf1936}. We have presented in Eqs.~(\\ref{sinE}), (\\ref{sinB}), (\\ref{hagensin}), (\\ref{pertur}) closed form results for the one-loop effective Lagrangian $\\Delta{\\mathcal L}_{\\rm eff}$ (\\ref{oneloopl}) for arbitrary strength of constant electromagnetic fields. The results will receive fluctuation corrections from higher loop diagrams. These carry one or more factors $ \\alpha $, $ \\alpha ^2$, \\dots~ and are thus suppressed by factors $1/137$. Thus results are valid for all field strengths with an error no larger than roughly 1\\%. If we include, for example, the two-loop correction, the first term in the Heisenberg and Euler effective Lagrangian (\\ref{Kleinert1}) becomes \\cite{Kleinert2008} % \\begin{equation} (\\Delta{\\mathcal L}_{\\rm eff})_{\\mathcal P} = \\frac{2 \\alpha}{90 \\pi E_c^2}\\left\\{ \\left(1+\\frac{40 \\alpha }{9\\pi}\\right)({\\bf E}^2\\!-\\!{\\bf B}^2)^2+7 \\left(1+\\frac{1315 \\alpha }{252\\pi} \\right)({\\bf E}\\cdot {\\bf B})^2 \\right\\}. \\label{Kleinert2} \\end{equation} % Readers can consult the recent review article \\cite{Dunne:2004nc}, where one finds discussions and computations of the effective Lagrangian at the two-loop level, and \\cite{2000NuPhB.564..591D} for discussion of pair production rate. \\subsection{Theory of pair production in an alternating field}\\label{alternating} When the external electromagnetic field $F^{\\rm e}_{\\mu\\nu}$ is space-time-dependent, i.e., $F^{\\rm e}_{\\mu\\nu}=F^{\\rm e}_{\\mu\\nu}({\\bf x},t)$ the exponential in Eq.~(\\ref{probability01}) can no longer be calculated exactly. In this case, JWKB methods have to be used to calculate pair production rates \\cite{1970PhRvD...2.1191B,1971ZhPmR..13..261P,1972JETP...34..709P,2001JETPL..74..133P,1972JETP...35..659P,1973JETPL..17..368M,Marinov1977,1971ZhPmR..13..261P,1972JETP...34..709P,2001JETPL..74..133P}. The aim of this section is to show how one can use a semi-classical JWKB approach to estimate the rate of pair production in an oscillating electric field as first indicated by Brezin and Itzykson in Ref.~\\cite{1970PhRvD...2.1191B}. They evaluated the production rate of charged boson pairs. The results they obtained can be straightforwardly generalized to charged fermion case, since the spins of charged particles contribute essentially with a counting factor to the final results (see Secs.~\\ref{semi} and \\ref{Schwingerformula}). Thus, let the electromagnetic potential be $\\hat {\\bf z}$ directed, uniform in space and periodic in time with frequency $\\omega_0$: \\begin{equation} A^{\\rm e}_\\mu(x)=(0,0,0,A(t)),\\quad A(t)=\\frac{E}{\\omega_0}\\cos\\omega_0 t. \\label{alterpotential} \\end{equation} Then the electric field is $\\hat {\\bf z}$ directed, uniform in space and periodic in time as well. The electric field strength is given by $E(t)=-\\dot A(t)=E\\sin\\omega_0 t$. It is assumed that the electric field is adiabatically switched on and damped off in a time $T^{\\rm e}$, which is much larger than the period of oscillation $T_0=2\\pi/\\omega_0$. Suppose also that $T_0$ is much larger than the Compton time $2\\pi/\\omega$ of the created particle , i.e., \\begin{equation} T^{\\rm e}\\gg T_0\\gg \\frac{2\\pi}{\\omega}\\simeq \\frac{2\\pi}{m_e}, \\label{thyrachy} \\end{equation} where $\\omega=\\sqrt{|{\\bf p}|^2+m_e^2}$, ${\\bf p}$ being the 3-momentum of the created particle. Furthermore, $eE$ is assumed to be much smaller than $m_e^2$, i.e., $E\\ll E_c$ (see Eq.~(\\ref{wkbc1})). We have to study the time evolution of a scattered wave function $\\psi(t)$ representing the production of particle and antiparticle pairs in the electromagnetic potential (\\ref{alterpotential}). As usual, an antiparticle can be thought of as a wave-packet moving backward in time. Therefore, for large positive time (forward) only positive energy modes ($\\sim e^{-i\\omega t}$) contribute to $\\psi(t)$. Similarly, for large negative times both positive energy and negative energy modes ($\\sim e^{i\\omega t}$) contribute to $\\psi(t)$ which satisfies the differential equation \\cite{1970PhRvD...2.1191B}: \\begin{equation} \\left[\\frac{d^2}{dt^2}+\\omega^2(t)\\right]\\psi(t)=0, \\label{bscatt} \\end{equation} where the ``variable frequency'' is defined as \\begin{equation} \\omega(t)\\equiv \\left\\{m_e^2+{\\bf p_\\perp}^2+[p_z-eA(t)]^2\\right\\}^{1/2}. \\label{bscattw} \\end{equation} The JWKB method suggests a general solution of the from \\begin{equation} \\psi(t)=\\alpha(t)e^{-i\\chi(t)}+\\beta(t)e^{i\\chi(t)}, \\quad \\chi(t)\\equiv \\int_0^tdt'\\omega(t'), \\label{bscattwave} \\end{equation} where the boundary conditions at large positive and negative times are: \\begin{equation} \\alpha(-\\infty)=1,\\quad \\beta(+\\infty)=0; \\quad \\dot\\chi(\\pm\\infty)=\\omega. \\label{bscattcond} \\end{equation} The backward scattering amplitude $\\beta(t)$ for large negative time ($t\\rightarrow -\\infty$) represents the probability of antiparticle production. The normalization condition $|\\psi(t)|^2=1$ implies \\begin{equation} \\dot\\alpha(t)e^{-i\\chi(t)}+\\dot\\beta(t)e^{i\\chi(t)}=0. \\end{equation} Eq.~(\\ref{bscatt}) can be written in terms of the scattering amplitudes as \\begin{equation} \\dot\\alpha(t)e^{-i\\chi(t)}-\\dot\\beta(t)e^{i\\chi(t)} =-\\frac{\\dot\\omega(t)}{\\omega(t)}\\left[\\alpha(t)e^{-i\\chi(t)}-\\beta(t)e^{i\\chi(t)}\\right], \\label{bscattwaven} \\end{equation} or, which is the same, \\begin{eqnarray} \\dot\\alpha(t)&=&-\\frac{\\dot\\omega(t)}{2\\omega(t)}\\left[\\alpha(t)-\\beta(t)e^{i2\\chi(t)}\\right], \\label{bscattwaven1}\\\\ \\dot\\beta(t)&=&-\\frac{\\dot\\omega(t)}{2\\omega(t)}\\left[\\beta(t)-\\alpha(t)e^{-i2\\chi(t)}\\right]. \\label{bscattwaven2} \\end{eqnarray} It follows from assumption (\\ref{thyrachy}) that $\\dot\\omega(t)$ vanishes as $|t|\\rightarrow\\infty$, i.e., \\begin{equation} \\frac{\\dot\\omega(t)}{\\omega^2(t)}=\\frac{eE[p_z-eA(t)]}{\\left\\{m_e^2+{\\bf p_\\perp}^2+[p_z-eA(t)]^2\\right\\}^{3/2}} \\ll 1. \\end{equation} More precisely \\begin{equation} \\left|\\frac{\\dot\\omega(t)}{\\omega^2(t)}\\right| <\\frac{eE}{m_e^2+{\\bf p_\\perp}^2}<\\frac{eE}{m_e^2}\\ll 1. \\label{bscattwr} \\end{equation} Therefore, $\\alpha(t)$ and $\\beta(t)$ slowly vary in time and tend to constants as $|t|\\rightarrow\\infty$. The phase $e^{i2\\chi(t)}$ oscillates very rapidly as compared to the variation of $\\alpha(t)$ and $\\beta(t)$, for $\\dot\\chi(t)=\\omega(t)\\gg |\\dot\\omega(t)/\\omega(t)|$. In the zeroth order the oscillating terms in Eqs.~(\\ref{bscattwaven1}), (\\ref{bscattwaven2}) are negligible and one finds \\begin{equation} \\alpha^{(0)}(t)=[\\omega/\\omega(t)]^{1/2}\\simeq 1;\\quad \\beta^{(0)}(t)=0, \\label{1iteration} \\end{equation} which duly satisfy the boundary conditions, and \\begin{equation} \\beta^{(1)}(t)=\\int_t^\\infty dt'\\frac{\\dot\\omega(t')}{2\\omega(t')}e^{-i2\\chi(t')}, \\label{bscattwavesolu} \\end{equation} where (\\ref{thyrachy}) and (\\ref{bscattwr}) have been used. $|\\beta^{(1)}(-\\infty)|^2$ gives information about the probability of particle-antiparticle pair production. Namely, the probability of pair production per unit volume and time is given by \\begin{eqnarray} \\tilde {\\mathcal P} &=& \\lim_{T^{\\rm e}\\rightarrow \\infty}\\frac{1}{T^{\\rm e}} \\int \\frac{d^3k}{(2\\pi)^3}|\\beta^{(1)}(-T^{\\rm e})|^2\\nonumber\\\\ &=&\\int \\frac{d^3k}{(2\\pi)^3}\\lim_{T^{\\rm e}\\rightarrow \\infty}\\frac{1}{T^{\\rm e}} \\left|\\int_{-T^{\\rm e}/2}^{T^{\\rm e}/2} dt'\\frac{\\dot\\omega(t')}{2\\omega(t')}e^{-i2\\chi(t')}\\right|^2. \\label{bscattpro} \\end{eqnarray} Since $\\omega(t)$ is a periodic function with the same frequency $\\omega_0$ as $A(t)$ one can make a Fourier series expansion: \\begin{equation} \\frac{\\dot\\omega(t)}{2\\omega(t)}=\\sum_{n=-\\infty}^{+\\infty}c_n e^{in\\omega_0}. \\label{fouriou} \\end{equation} Defined a renormalized frequency $\\Omega$ via $\\chi(t)=t\\Omega$ one finds \\begin{equation} \\Omega\\equiv \\int^{2\\pi}_0\\frac{dx}{2\\pi}\\left[m_e^2+{\\bf k_\\perp}^2+ \\left(k_3-\\frac{eE}{\\omega_0}\\cos x\\right)^2\\right]^{1/2}. \\label{norw} \\end{equation} so that \\begin{equation} \\lim_{T^{\\rm e}\\rightarrow \\infty}\\frac{1}{T^{\\rm e}} \\left|\\int_{-T^{\\rm e}/2}^{T^{\\rm e}/2} dt'\\frac{\\dot\\omega(t')}{2\\omega(t')}e^{-i2\\chi(t')}\\right|^2 =2\\pi\\sum_n\\delta(n\\omega_0-2\\Omega)|c_n|^2. \\label{bscattdel} \\end{equation} Consequently, the probability of pair production (\\ref{bscattpro}) is, \\begin{equation} \\tilde {\\mathcal P}=\\int \\frac{d^3k}{(2\\pi)^2} \\sum_n\\delta(n\\omega_0-2\\Omega)|c_n|^2=\\int \\frac{d^3k}{(2\\pi)^2\\omega_0} |c_{n^\\circ}|^2, \\label{bscattpro1} \\end{equation} where ${n^\\circ}=2\\Omega/\\omega_0$ and $c_{n^\\circ}$ are determined via Eq.~(\\ref{fouriou}) as \\begin{equation} c_{n^\\circ}=\\int_{-\\pi}^{\\pi} \\frac{dx}{2\\pi} \\frac{\\dot\\omega(x)}{2\\omega(x)}\\exp\\left\\{\\frac{2i}{\\omega_0}\\int^x_0dx' \\left[m_e^2+{\\bf p_\\perp}^2+\\left(p_z-\\frac{eE}{\\omega_0}\\cos(x')\\right)^2\\right]^{1/2} \\right\\}. \\label{bscattpro2} \\end{equation} The expression for $c_{n^\\circ}$ contains a very rapidly oscillating phase factor with frequency of the order of $m_e/\\omega_0$, and it decreases very rapidly in terms of imaginary time $\\tau=-it$. Its evaluation requires the application of the steepest-descent method in the complex time $x=\\omega_0 t$ plane. This is done by selecting a proper contour turning in a neighborhood of the saddle point and following the steepest-descent line, so as to find the main contributions to the integral in Eq.~(\\ref{bscattpro2}). The saddle point originates from branch points and poles in Eq.~(\\ref{bscattpro2}), which are the zeros of $\\omega(x)$. Mathematical details can be found in Ref.~\\cite{1970PhRvD...2.1191B}. One finds \\begin{equation} \\tilde {\\mathcal P} \\simeq\\frac{\\omega_0}{9}\\int \\frac{d^3k}{(2\\pi)^2} e^{-2A}\\cos^2B, \\label{bscattpro3} \\end{equation} where \\begin{equation} -A+iB=\\frac{2i}{\\omega_0}\\int^{x_0}_0dx' \\left[m_e^2+{\\bf p_\\perp}^2+\\left(p_z-\\frac{eE}{\\omega_0}\\cos(x')\\right)^2\\right]^{1/2}, \\label{complexab} \\end{equation} and the saddle point is $x_0=1/\\pi+i\\sinh^{-1}[(\\omega_0/eE)(m_e^2+{\\bf p_\\perp}^2)^{1/2}]$. The exponential factor $e^{-2A}$ in Eq.~(\\ref{bscattpro3}) indicates that particle-antiparticle pairs tend to be emitted with small momenta. This allows one to estimate the right-hand side of Eq.~(\\ref{bscattpro3}) as follow: (i) $p_z$ is set equal to zero, moreover, the range of the $p_z$-integration is of the order of $2eE/\\omega_0$ as suggested by the classical equation of motion (\\ref{bscatt}); (ii) $\\cos^2B$ is replaced by its average value $1/2$. As a result, one obtains \\cite{1970PhRvD...2.1191B}, \\begin{equation} \\tilde {\\mathcal P}\\simeq\\frac{(eE)^3}{18\\pi\\omega_0^2}\\int_{\\eta^{-1}}^\\infty du u \\exp\\left[-\\frac{\\pi eE}{\\omega_0^2}u^2g(u)\\right], \\label{bscattpro4} \\end{equation} where $\\eta^{-1}=m_e\\omega_0/(eE)$, $u=(m_e^2+{\\bf p_\\perp}^2)\\omega_0^2/(eE)^2$ and \\begin{equation} g(z) = \\frac{4}{\\pi}\\int_0^1dy \\Big[\\frac{1-y^2}{ 1 + z^{-2} y^2}\\Big]^{1/2}=F\\left(\\frac{1}{2},\\frac{1}{2};2;-z^{-2}\\right), \\label{xlasereta} \\end{equation} where $F(1/2,1/2;2;-z^{-2})={}_2F_1(1/2,1/2;2;-z^{-2})$ is the Gauss hypergeometrical function. The function $u^2g(u)$ is monotonically increasing: \\begin{equation} \\frac{eE}{\\omega_0^2}u^2g(u)\\ge \\frac{eE}{\\omega_0^2}\\eta^{-2}g(\\eta) = \\frac{m_e^2}{eE}g(\\eta)\\gg 1, \\label{intlimit} \\end{equation} which indicates that the integral in (\\ref{bscattpro4}) is strongly dominated by values in a neighborhood of $u=\\eta^{-1}$. This allows one to approximately perform the integration and leads to the rate of pair production of charged bosons \\cite{1970PhRvD...2.1191B}, \\begin{equation} \\tilde {\\mathcal P}_{\\rm boson}\\simeq\\frac{\\alpha E^2}{2\\pi}\\frac{1}{g(\\eta)+\\frac{1}{2\\eta} g'(\\eta)} \\exp\\left[-\\frac{\\pi m_e^2}{eE}g(\\eta)\\right]. \\label{bscattpro5} \\end{equation} Analogously, the rate of pair production of charged fermions can be approximately obtained from Eq.~(\\ref{bscattpro5}) by taking into account two helicity states of fermions (see Secs.~\\ref{semi} and \\ref{Schwingerformula}), \\begin{equation} \\tilde {\\mathcal P}_{\\rm fermion}\\simeq\\frac{\\alpha E^2}{\\pi}\\frac{1}{g(\\eta)+\\frac{1}{2\\eta} g'(\\eta)} \\exp\\left[-\\frac{\\pi m_e^2}{eE}g(\\eta)\\right]. \\label{bscattpro5f} \\end{equation} This formula has played an important role in recent studies of electron and positron pair production by laser beams, which we will discuss in some details in Section~\\ref{Xray}. Momentum spectrum of electrons and positrons, produced from the vacuum, was calculated in \\cite{1971ZhPmR..13..261P,1972JETP...34..709P,1972JETP...35..659P,2001JETPL..74..133P}. For $\\eta\\gg1$ this distribution is concentrated along the direction of electric field, while for $\\eta\\ll1$ it approaches isotropic one. Unfortunately, it appears very difficult to produce a macroscopic electric field with strength of the order of the critical value (\\ref{critical1}) and lifetime long enough ($\\gg \\hbar/(m_ec^2)$) in any ground laboratory to directly observe the Sauter-Euler-Heisenberg-Schwinger process of electron--positron pair production in vacuum. The same argument applies for the production of any other pair of fermions or bosons. In the following Section, we discuss some ideas to experimentally create a transient electric field $E\\lesssim E_c$ in Earth-bound laboratories, whose lifetime is expected to be long enough (larger than $\\hbar/m_ec^2$) for the pair production process to take place. \\subsection{Nonlinear Compton scattering and Breit-Wheeler process} In Section~\\ref{BW}, we have discussed the Breit--Wheeler process \\cite{1934PhRv...46.1087B} in which an electron--positron pair is produced in the collision of two real photons $\\gamma_1 + \\gamma_2 \\rightarrow e^+ +e^-$ (\\ref{2gammaee}). The cross-section they obtained is $O(r_e^2)$, where $r_e$ is the classical electron radius, see Eq.~(\\ref{bwsection3}). This lowest order photon-photon pair production cross-section is so small that it is difficult to observe creation of pairs in the collision of two high-energy photon beams, even if their center of mass energy is larger than the energy-threshold $2m_ec^2= 1.02$ MeV. In the previous Sections we have seen that in strong electromagnetic fields in lasers the effective nonlinear terms (\\ref{Kleinert1}) become significant and therefore, the interaction needs not to be limited to initial states of two photons \\cite{1962JMP.....3...59R, 1971PhRvL..26.1072R}. A collective state of many interacting laser photons occurs. We turn now to two important processes \\cite{1996PhRvL..76.3116B, 1997PhRvL..79.1626B} emerging in the interaction of an ultrarelativistic electron beam with a terawatt laser pulse, performed at SLAC \\cite{1996NIMPA.383..309K}, when strong electromagnetic fields are involved. The first process is the nonlinear Compton scattering, in which an ultrarelativistic electron absorbs multiple photons from the laser field, but emits only a single photon via the process \\begin{equation} e+n\\omega \\rightarrow e' + \\gamma , \\label{process1} \\end{equation} where $\\omega$ represents photons from the strong electromagnetic wave of the laser beam (its frequency being $\\omega$), $n$ indicates the number of absorbed photons and $\\gamma$ represents a high-energy emitted photon (see Eq.~(\\ref{process2}) for {\\it cross symmetry}). The theory of this nonlinear Compton effect (\\ref{process1}) is given in Section \\ref{lighttheoryquantum}. The same process (\\ref{process1}) has been expressed by Bamber et al. \\cite{1999PhRvD..60i2004B} in a semi-classical framework. The second is the nonlinear Breit--Wheeler process \\begin{equation} \\gamma+n\\omega \\rightarrow e^+ +e^- . \\label{process2} \\end{equation} between this very high-energy photon $\\gamma$ and multiple laser photons: the high-energy photon $\\gamma$, created in the first process, propagates through the laser field and interacts with laser photons $n\\omega$ to produce an electron--positron pair \\cite{1997PhRvL..79.1626B}. In the electric field $E$ of an intense laser beam, an electron oscillates with the frequency $\\omega$ of the laser and its maximum velocity in unit of the speed of light is given by \\begin{equation} v_{\\rm max}\\gamma_{\\rm max}=\\frac{eE}{m\\omega},\\quad\\quad \\gamma_{\\rm max}=1/\\sqrt{1-v_{\\rm max}^2}. \\label{maxv} \\end{equation} In the case of weak electric field, $v_{\\rm max}\\ll 1$ and the nonrelativistic electron emits the {\\it dipole} radiation well described in linear and perturbative QED. On the other hand, in the case of strong electric fields, $v_{\\rm max}\\rightarrow 1$ and the ultrarelativistically oscillating electron emits {\\it multi-pole} radiation. The radiated power is then a nonlinear function of the intensity of the incident laser beam. Using the maximum velocity $v_{\\rm max}$ of oscillating electrons in the electric field of laser beam, one can characterize the effect of nonlinear Compton scattering by the dimensionless parameter \\begin{equation} \\eta =v_{\\rm max}\\gamma_{\\rm max}=\\frac{eE_{\\rm rms}}{ m\\omega }=\\frac{m_ec^2}{ \\omega\\hbar}\\frac{E_{\\rm rms}}{ E_c}, \\label{eta} \\end{equation} where the subscript `rms' means root-mean-square, with respect to the number of interacting laser photons with scattered electron. The parameter $\\eta$ can be expressed as a Lorentz invariant, \\begin{equation} \\eta^2=\\frac{e^2|\\langle A_\\mu A^\\mu \\rangle|}{ m_e^2}, \\label{eta1} \\end{equation} where $A_\\mu$ is the gauge potential of laser wave, $\\partial^\\mu A_\\mu=0$ and the time-average is taken over one period of laser wave, $\\langle A_\\mu \\rangle =0$ and \\begin{equation} \\langle A_\\mu A^\\mu \\rangle=\\langle (A_\\mu -\\langle A_\\mu \\rangle)^2\\rangle. \\label{aaverage} \\end{equation} Eq.~(\\ref{eta1}) shows that $\\eta^2$ is the intensity parameter of laser fields, and $\\eta$ in (\\ref{eta}) coincides with the parameter $\\eta$ introduced in Eq.~(\\ref{bscattpro4}) for the pair production in an alternating electric field (see Section~\\ref{alternating}). %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \\subsection{Quantum description of nonlinear Compton effect}\\label{lighttheoryquantum} In Refs.~\\cite{1962JMP.....3...59R, 1971PhRvL..26.1072R, Nikishov1964, Nikishov1964a, Nikishov1965, 1967JETP...25.1135N, Nikishov1979, Nikishov1965a, Sengupta1952, 1964PhRv..133..705B, Goldman1964, 1964PhL.....8..103G, Eberly1969, 1982els..book.....B}, the quantum theory of the interaction of free electrons with the field of a strong electromagnetic wave has been studied. The application of quantum perturbation theory to such interaction requires not only that the interaction constant $\\alpha$ should be small but also that field should be sufficiently weak. The characteristic quantity in this respect is the dimensionless invariant ratio $\\eta$, see (\\ref{eta1}). The photon emission processes occurring in the interaction of an electron with the field of a strong electromagnetic wave have been discussed in Ref.~\\cite{1982els..book.....B} for any $\\eta$ value. The method used is based on an exact treatment of this interaction, while the interaction of the electron with the newly emitted photons regarded as a perturbation. Laser beam is considered as a monochromatic plane wave, described by the gauge potential $A_\\mu(\\phi)$ and $\\phi=kx$, where wave vector $k=(\\omega, {\\bf k})$ $(k^2=0)$ (see Eq.~(\\ref{2p2epotential})). The Dirac equation can be exactly solved \\cite{Wolkow1935} for an electron moving in this field of electromagnetic plane wave of an arbitrary polarization and the normalized wave function of the electron with momentum $p$ is given by (c.e.g. \\cite{1982els..book.....B}), \\begin{eqnarray} \\psi_p &=& \\left[1+\\frac{e}{2(kp)}\\not\\!{k}\\not{\\hspace{-5pt}A}\\right]\\frac{u(p)}{\\sqrt{2q_0}}e^{i\\Phi},\\label{eplanes}\\\\ \\Phi &=& -px-\\int^{kx}_0\\left[\\frac{e}{(kp)}(pA)-\\frac{e^2}{2(kp)}A^2\\right]d\\phi, \\label{ephase} \\end{eqnarray} where $u(p)$ is the solution of free Dirac equation $(\\not\\!{p}-m_e)u(p)=0$ and the time-average value of 4-vector, \\begin{equation} q= p -\\frac{e^2\\langle A^2\\rangle}{2(kp)}k , \\label{tqmomen} \\end{equation} is the kinetic momentum operator in the electron state $\\psi_p$ (\\ref{eplanes}) and the ``effective mass'' $m_*$ of the electron in the field is \\begin{equation} q^2=m_*^2, \\quad m_*=m_e\\sqrt{1+\\eta^2 }, \\label{effectivemass} \\end{equation} where $\\eta^2$ is given by (\\ref{eta1}). The electron becomes ``heavy'' in an oscillating electromagnetic field. The $S$-matrix element for a transition of the electron from the state $\\psi_p$ to the state $\\psi_{p'}$, with emission of a photon having momentum $k'$ and polarization $\\epsilon'$ is given by (c.e.g. \\cite{1982els..book.....B}) \\begin{eqnarray} S_{fi}&=&-ie\\int \\bar\\psi_{p'}(\\gamma \\epsilon'{}^*)\\psi_p\\frac{e^{ik'x}}{\\sqrt{2\\omega'}}d^4x\\label{eplaneS1}\\\\ &=&\\frac{1}{2\\omega'\\cdot 2q_0\\cdot 2q_0'}\\sum_n M^{(n)}_{fi}(2\\pi)^4i\\delta^{(4)}(nk+q-q'-k'), \\label{eplaneS} \\end{eqnarray} where the integrand in Eq.~(\\ref{eplaneS1}) is expanded in Fourier series and expansion coefficients are in terms of Bessel functions $J_n$, the scattering amplitude $M^{(n)}_{fi}$ in Eq.~(\\ref{eplaneS}) is obtained\\footnote{The explicit expression $M^{(n)}_{fi}$ is not given here for its complexity, see for example \\cite{1982els..book.....B}.} by integrating over $x$. Eq.~(\\ref{eplaneS}) shows that $S_{fi}$ is an infinite sum of terms, each corresponds to an energy-momentum conservation law $nk+q=q'+k'$, indicating an electron ($q$) absorbs $n$-photons ($nk$) and emits another photon ($k'$) of frequency \\begin{equation} \\omega'=\\frac{n\\omega}{1+(n\\omega/m_*)(1-\\cos\\theta)}, \\label{gammaenergy} \\end{equation} in the frame of reference where the electron is at rest (${\\bf q}=0,q_0=m_*$), and $\\theta$ is the angle between $\\bf k$ and $\\bf k'$. Given the $n$th term of the $S$-matrix $S_{fi}$ (\\ref{eplaneS}), the differential probability per unit volume and unit time yields, \\begin{equation} d{\\mathcal P}_{e\\gamma}^{(n)}= \\frac{d^3{\\bf k}'d^3{\\bf q}'}{(2\\pi)^6\\cdot 2\\omega'\\cdot 2q_0\\cdot 2q_0'}|M^{(n)}_{fi}|^2(2\\pi)^4i\\delta^{(4)}(nk+q-q'-k'). \\label{eplaneP} \\end{equation} Integrating over the phase space of final states $\\int d^3{\\bf k}'d^3{\\bf q}'$, one obtains the total probability of emission from unit volume in unit time (circular polarization), \\begin{equation} {\\mathcal P}_{e\\gamma}=\\frac{e^2m_e^2}{4q_0} \\sum_{n=1}^\\infty\\int_0^{\\kappa_n}\\frac{d\\kappa}{(1+\\kappa)^2} \\left[-4J^2_n(z)+\\eta^2\\big(2+\\frac{\\kappa^2}{1+\\kappa}\\big) (J^2_{n+1}+J^2_{n-1}-2J^2_n) \\right], \\label{epprobability} \\end{equation} where $\\kappa=(kk')/(kp')$, $\\kappa_n=2n(kp)/m_*^2$ and Bessel functions $J_n(z)$, \\begin{equation} z=2m_e^2\\frac{\\eta}{(1+\\eta^2)^{1/2}} \\left[\\frac{\\kappa}{\\kappa_n} \\left(1-\\frac{\\kappa}{\\kappa_n}\\right)\\right]^{1/2}, \\label{bargument} \\end{equation} for any $\\eta$ value. A systematic investigation of various quantum processes in the field of a strong electromagnetic wave can be found in \\cite{Nikishov1964, Nikishov1964a, Nikishov1965, 1967JETP...25.1135N, Nikishov1979, Nikishov1965a}, in particular photon emission and pair production in the field of a plane wave with various polarizations are discussed. We now turn to the Breit--Wheeler process for multi-photons (\\ref{process2}). In this process, the pair production is attributed to the interaction of a high-energy photon with many laser photons in the electromagnetic laser wave. Actually, the Breit--Wheeler process for multi-photons, see Eq. (\\ref{process2}), is related to the nonlinear Compton scattering process, see Eq. (\\ref{process1}), by {\\it crossing symmetry}. By replacement $p\\rightarrow -p$ and $k'\\rightarrow -l$ and reverse the common sign of the expression in Eq.~(\\ref{eplaneS}), one obtains the probability of pair production (\\ref{process2}) by a photon $\\gamma$ (momentum $l$) colliding with $n$ laser photons (momentum $k$) per unit volume in unit time (circular polarization) \\cite{Nikishov1964, Nikishov1964a, Nikishov1965, 1967JETP...25.1135N, Nikishov1979, Nikishov1965a}, \\begin{eqnarray} {\\mathcal P}_{\\gamma\\gamma}&=&\\frac{e^2m_e^2}{16l_0}\\sum_{n>n_0}^\\infty\\int_1^{\\upsilon_n} \\frac{d\\upsilon}{\\upsilon^{3/2}(1+\\upsilon)^{1/2}}\\times\\nonumber\\\\ &\\times&\\left[2J^2_n(z)+\\eta^2(2\\upsilon-1)(J^2_{n+1}+J^2_{n-1}-2J^2_n) \\right], \\label{ggprobability} \\end{eqnarray} where $\\upsilon=(kl)^2/4(kq)(kq')$, $\\upsilon_n=n/n_0$, $n_0=2m_*^2/(kl)$ and Bessel functions $J_n(z)$, \\begin{equation} z=4m_e^2\\frac{\\eta (1+\\eta^2)^{1/2}}{(kl)} \\left[\\frac{\\upsilon}{\\upsilon_n}\\left(1-\\frac{\\upsilon}{\\upsilon_n}\\right)\\right]^{1/2}. \\nonumber \\end{equation} In Eq.~(\\ref{ggprobability}), the number $n$ of laser photons must be larger than $n_0$ ($n>n_0$), which is the energy threshold $n_0(kl)=2m_*^2$ for the process (\\ref{process2}) of pair production to occur. ", "conclusions": "\\label{remarks} We have reviewed three fundamental quantum processes which have highlighted some of the greatest effort in theoretical and experimental physics in last seventy years. They all deal with creation and annihilation of electron--positron pairs. We have followed the original path starting from the classical works of Dirac, on the process $e^+e^-\\rightarrow 2\\gamma$, and the inverse process, $2\\gamma\\rightarrow e^+e^-$, considered by Breit--Wheeler. We have then reviewed the $e^+e^-$ pair creation in a critical electric field $E_c=m_e^2c^3/(\\hbar e)$ and the Sauter-Heisenberg-Euler-Schwinger description of this process both in Quantum Mechanics and Quantum Electro-Dynamics. We have also taken this occasion to reconstruct the exciting conceptual developments, initiated by the Sauter work, enlarged by the Born-Infeld nonlinear electrodynamical approach, finally leading to the Euler and Euler-Heisenberg results. We were guided in this reconstruction by the memories of many discussions of one of us (RR) with Werner Heisenberg. We have then reviewed the latest theoretical developments deriving the general formula for pair production rate in electric fields varying in space and in time, compared with one in a constant electric field approximation originally studied by Schwinger within QED. We also reviewed recent studies of pair production rates in selected electric fields varying both in space and in time, obtained in the literature using instanton and JWKB methods. Special attention has been given to review the pair production rate in electric fields alternating periodically in time, early derived by Brezin, Itzykson and Popov, and the nonlinear Compton effect in the processes of electrons and photons colliding with laser beams, studied by Nikishov and Narozhny. These theoretical results play an essential role in Laboratory experiments to observe the pair production phenomenon using laser technologies. We then reviewed the different level of verification of these three processes in experiments carried all over the world. We stressed the success of experimental verification of the Dirac process, by far one of the most prolific and best tested process in the field of physics. We also recalled the study of the hadronic branch in addition to the pure electrodynamical branch originally studied by Dirac, made possible by the introduction of $e^+e^-$ storage rings technology. We then turned to the very exciting current situation which sees possibly the Breit--Wheeler formula reaching its first experimental verification. This result is made possible thanks to the current great developments of laser physics. We reviewed as well the somewhat traumatic situation in the last forty years of the heavy-ion collisions in Darmstadt and Brookhaven, yet unsuccessfully attempting to observe the creation of electron--positron pairs. We also reviewed how this vast experimental program was rooted in the theoretical ideas of Zeldovich, Popov, Greiner and their schools. We have also recalled the novelty in the field of relativistic astrophysics where we are daily observing the phenomenon of Gamma Ray Bursts \\cite{1999PhR...314..575P, 2005RvMP...76.1143P, 2006RPPh...69.2259M, Ruffini2003, 2007AIPC..910...55R}. These bursts of photons occur in energy range keV to MeV, last about one second and come from astrophysical sources located at a cosmological distance \\cite{1997Natur.387..783C,1997Natur.386..686V, 1998Natur.393...35K, 1998Natur.393...41H, 1998Natur.393...43R}. The energy released is up to $\\sim 10^{55}$ ergs, equivalent to all the energy emitted by all the stars of all the galaxies of the entire visible Universe during that second. It is generally agreed that the energetics of these GRB sources is dominated by a dense plasma of electrons, positrons and photons created during the process of gravitational collapse leading to a Black Hole, see e.g. \\cite{1975PhRvL..35..463D,1999A&A...350..334R,2000A&A...359..855R,1998A&A...338L..87P} and references therein. The Sauter-Heisenberg-Euler-Schwinger vacuum polarization process, we have considered in the first part of the report, is a classic theoretical model to study the creation of an electron--positron optically thick plasma. Similarly the Breit--Wheeler and the Dirac processes we have discussed, are essential in describing the further evolution of such an optically thick electron--positron plasma. The GRBs present an unique opportunity to test new unexplored regime of ultrahigh energy physics with Lorentz factor $\\gamma\\sim 100-1000$ and relativistic field theories in the strongest general relativistic domain. The aim in this report, in addition to describe the above mentioned three basic quantum processes, has been to identify and review three basic relativistic regimes dealing with an optically thin and optically thick electron--positron plasma. The first topic contains the basic results of the physics of black holes, of their energetics and of the associated process of vacuum polarization. We reviewed the procedures to generalize in a Kerr--Newman geometry the QED treatment of Schwinger and the creation of enormous number of $10^{60}$ electron--positron pairs in such a process. The second topic is the back reaction of a newly created electron--positron plasma on an overcritical electric field. Again we reviewed the Breit--Wheeler and Dirac processes applied in the wider context of the Vlasov--Boltzmann--Maxwell equations. To discuss the back reaction of electron--positron pair on external electric fields, we reviewed semi-classical and kinetic theories describing the plasma oscillations using respectively the Dirac-Maxwell equations and the Boltzmann--Vlasov equations. We also reviewed the discussions of plasma oscillations damping due to quantum decoherence and collisions, described by respectively the quantum Boltzmann--Vlasov equation and Boltzmann--Vlasov equation with particle collisions terms. We particularly addressed the study of the influence of the collision processes $e^{+}e^{-}\\rightleftarrows \\gamma\\gamma$ on the plasma oscillations in supercritical electric field $E > E_c$. After $10^{3-4}$ Compton times, the oscillating electric field is damped to its critical value with a large number of photons created. An equipartition of number and energy between electron--positron pairs and photons is reached. For the plasma oscillation with undercritical electric field $E \\lesssim E_c$, we recalled that electron--positron pairs, created by the vacuum polarization process, move as charged particles in external electric field reaching a maximum Lorentz factor at finite length of oscillations, instead of arbitrary large Lorentz factors, as traditionally assumed. Finally we point out some recent results which differentiate the case $E>E_{c}$ from the one $EE_{c}$ the vacuum polarization process transforms the electromagnetic energy of the field mainly in the rest mass of pairs, with moderate contribution to their kinetic energy. Such phenomena, certainly fundamental on astrophysical scales, may become soon directly testable in the verification of the Breit--Wheeler process tested in laser experiments in the laboratory. As the third topic we have reviewed the recent progress in the understanding of thermalization process of an optically thick electron--positron-photon plasma. Numerical integration of relativistic Boltzmann equation with collisional integrals for binary and triple interactions is used to follow the time evolution of such a plasma, in the range of energies per particle between 0.1 and 10 MeV, starting from arbitrary nonequilibrium configuration. It is recalled that there exist two types of equilibria in such a plasma: kinetic equilibrium, when all particles are at the same temperature, but have different nonzero chemical potential of photons, and thermal equilibrium, when chemical potentials vanish. The crucial role of direct and inverse binary and triple interactions in reaching thermal equilibrium is emphasized. In a forthcoming report we will address how the above mentioned three relativistic processes can be applied to a variety of astrophysical systems including neutron stars formation and gravitational collapse, supernovae explosions and GRBs. This report is dedicated to the progress of theoretical physics in extreme regimes of relativistic field theories which are on the verge of finding their experimental and observational verification in physics and astrophysics. It is then possible from our review and the many references we have given to gain a basic understanding of this new field of research. The three topics which we have reviewed are closely linked to the three quantum processes currently being tested in precision measurement in the laboratories. The experiments in the laboratories and the astrophysical observations cover complementary aspects which may facilitate a deeper and wider understanding of the nuclear and laser physics processes, of heavy-ion collisions as well as neutron stars formation and gravitational collapse, supernovae and GRBs phenomena. We shall return on such an astrophysical and observational topics in a dedicated forthcoming report. \\vskip1cm \\begin{center} *\\hskip3cm *\\hskip3cm * \\end{center} \\vskip1cm We are witnessing in these times some enormous experimental and observational successes which are going to be the natural ground to test some of the theoretical works which we have reviewed in this report. Among the many experimental progresses being done in particle accelerators worldwide we like to give special mention to two outstanding experimental facilities which are expected to give results in the forthcoming years. We refer here to the National Ignition Facility at the Lawrence Livermore National Laboratory to be soon becoming operational, see e.g. \\cite{2009Sci...324..326C} as well as the corresponding facility in France, the Mega Joule project \\cite{2006JPhy4.133..631G}. In astrophysics these results will be tested in galactic and extragalactic black holes observed in binary X-ray sources, active galactic nuclei, microquasars and in the process of gravitational collapse to a neutron star and also of two neutron stars to a black hole in GRBs. The progress there is equally remarkable. In the last few days after the completion of this report thanks to the tremendous progress in observational technology for the first time a massive hypergiant star has been identified as the progenitor of the supernova SN 2005gl \\cite{Gal-Yam2009}. In parallel the joint success of observations of the flotilla of X-ray observatories and ground-based large telescopes \\cite{Ruffini2009} have allowed to identify the first object ever observed at $z\\approx8$ the GRB090423 \\cite{2009arXiv0905.0001P}. To follow the progress of this field we are planning a new report which will be directed to the astrophysical nature of the progenitors and the initial physical conditions leading to the process of the gravitational collapse. There the electrodynamical structure of neutron stars, the phenomenon of the supernova explosion as well as theories of Gamma-Ray Bursts (GRBs) will be discussed. Both in the case of neutron stars and the case of black holes there are fundamental issues still to be understood about the process of gravitational collapse especially with the electrodynamical conditions at the onset of the process. The major difficulties appear to be connected with the fact that all fundamental interactions, the gravitational, the electromagnetic, the strong, the weak interactions appear to participate in essential way to this process which appear to be therefore one of the most interesting fundamental process of theoretical physics. Current progress is presented in the following works \\cite{2007IJMPD..16....1R, 2008AIPC..966..147R, 2008arXiv0804.3197R, 2008AIPC.1053..243R, 2008APS..APR8HE093R, 2008APS..APR8HE095R, 2009arXiv0903.3727P, 2009arXiv0903.4095R, 2009AIPC..........P, 2009AIPC..........R, 2009PhRvD........R, 2009PhRvDa.......R}. What is important to recall at this stage is only that both the supernovae and GRBs processes are among the most energetic and transient phenomena ever observed in the Universe: a supernova can reach energy of $\\sim 10^{52}$ ergs (hypernovae) on a time scale of a few months and GRBs can have emission of up to $\\sim 10^{55}$ ergs \\cite{2009Sci...323.1688A} in a time scale as short as of a few seconds. The central role in their description of neutron stars, for supernova as well as of black holes and the electron--positron plasma discussed in this report, for GRBs, are widely recognized. The reason which makes this last research so important can be seen in historical prospective: the Sun has been the arena to understand the thermonuclear evolution of stars \\cite{1968QB464.B46......}, Cyg X-1 has evidenced the gravitational energy role in explaining an astrophysical system \\cite{2003RvMP...75..995G}, the GRBs are promising to prove the existence for the first time of the ``blackholic energy''. These three quantum processes described in our report reveal the basic phenomena in the process of gravitational collapse predicted by Einstein theory of General Relativity \\cite{Ruffini2009}. \\vspace{0.3in}" }, "0910/0910.0699_arXiv.txt": { "abstract": "{Accurate transition probabilities for forbidden lines are important diagnostic parameters for low-density astrophysical plasmas. In this paper we present experimental atomic data for forbidden [\\ion{Fe}{ii}] transitions that are observed as strong features in astrophysical spectra. } {To measure lifetimes for the $3d^6$($^3$G)$4s$\\,a\\,$^4$G$_{11/2}$ and $3d^6$($^3$D)$4s$\\,b\\,$^4$D$_{1/2}$ metastable levels in \\ion{Fe}{ii} and experimental transition probabilities for the forbidden transitions $3d^7$\\,a\\,$^4$F$_{7/2,9/2}$ -- $3d^6$($^3$G)$4s$\\,a\\,$^4$G$_{11/2}$.} {The lifetimes were measured at the ion storage ring facility CRYRING using a laser probing technique. Astrophysical branching fractions were obtained from spectra of Eta Carinae, obtained with the Space Telescope Imaging Spectrograph onboard the {\\it Hubble Space Telescope}. The lifetimes and branching fractions were combined to yield absolute transition probabilities.} {The lifetimes of the a\\,$^4$G$_{11/2}$ and the b\\,$^4$D$_{1/2}$ levels have been measured and have the following values, $\\tau=0.75\\pm0.10$\\,s and $\\tau=0.54\\pm0.03$\\,s respectively. Furthermore, we have determined the transition probabilities for two forbidden transitions of a\\,$^4$F$_{7/2,9/2}$ -- a\\,$^4$G$_{11/2}$ at 4243.97 and 4346.85\\,\\AA . Both the lifetimes and the transition probabilities are compared to calculated values in the literature.} {} ", "introduction": "The cosmic abundance of iron is relatively high compared to other iron group elements and the spectrum of singly ionized iron, \\ion{Fe}{ii}, is a significant contributor to the spectral opacity of the sun and hotter stars. The complex energy level structure of \\ion{Fe}{ii} makes the spectrum extremely line rich and it has been studied in great detail with more than 1000 energy levels identified in the literature (\\citeauthor{Johansson09}\\,\\citeyear{Johansson09}). \\ion{Fe}{ii} lines are observed in the spectra of a wide variety of astronomical objects, and there is a considerable demand for accurate atomic data for this ion. To meet the accurate data requirements of modern astrophysics, a program was initiated to supply the astronomical community with reliable atomic data: The FERRUM-project (\\citeauthor{FERRUMproj}\\,\\citeyear{FERRUMproj}). The aim of this international collaboration is to measure and evaluate astrophysically relevant experimental and theoretical transition data for the iron group elements. There are 62 metastable levels in \\ion{Fe}{ii}. The parity forbidden lines from some of these levels are observed as prominent features in astrophysical low density plasmas, such as nebulae, \\ion{H}{ii} regions and circumstellar gas clouds. However, metastable levels have radiative lifetimes several orders of magnitude longer than other levels and are thus more affected by collisions. Due to the absence of these lines in laboratory spectra, the majority of forbidden line transition probabilities ($A$-values) available in the literature are from theoretical calculations. There are only four metastable levels in \\ion{Fe}{ii} with laboratory measured lifetimes. The a\\,$^6$S$_{5/2}$ and b\\,$^4$D$_{7/2}$ levels have been measured by \\citeauthor{Rostohar}\\,(\\citeyear{Rostohar}) using laser probing of a stored ion beam (a laser probing technique, LPT). In addition, the a\\,$^4$G$_{9/2}$ and b\\,$^2$H$_{11/2}$ levels have been measured by \\citeauthor{Hartman}\\,(\\citeyear{Hartman}) using the LPT. There is good agreement between \\citeauthor{Rostohar}\\,(\\citeyear{Rostohar}) and the calculated values of \\citeauthor{Nussbaumer}\\,(\\citeyear{Nussbaumer}) and \\citeauthor{Quinet}\\,(\\citeyear{Quinet}). However, the lifetimes of \\citeauthor{Rostohar}\\,(\\citeyear{Rostohar}) are systematically shorter than the calculations of Garstang (\\citeyear{Garstang}). \\citeauthor{Hartman}\\,(\\citeyear{Hartman}) also combined the lifetimes of a\\,$^6$S$_{5/2}$, b\\,$^4$D$_{7/2}$ and a\\,$^4$G$_{9/2}$ with branching fractions ($BF$s) to determine experimental $A$-values for forbidden transitions. \\citeauthor{Hartman}\\,\\citeyear{Hartman} measured the $BF$s in astrophysical spectra observed in the ejecta of Eta Carinae and presented additional theoretical $A$-value calculations. We present radiative lifetimes for the a\\,$^4$G$_{11/2}$ and b\\,$^4$D$_{1/2}$ metastable levels in \\ion{Fe}{ii} measured using the LPT at the CRYRING facility. In addition, $BF$s for two forbidden transitions 4243.97 \\AA\\ and 4346.85 \\AA\\ (a$^4$F$_{9/2}$ - a$^4$G$_{11/2}$ and a$^4$F$_{7/2}$ - a$^4$G$_{11/2}$) have been measured in astrophysical spectra observed in the ejecta of Eta Carinae recorded with the {\\it Hubble Space Telescope} ({\\it HST}) Space Telescope Imaging Spectrograph (STIS). The radiative lifetimes have been combined with the $BF$s to yield $A$-values and we provide a comparison with theoretical values in the literature. ", "conclusions": "The relativistic Hartree-Fock (HFR) calculations by \\citeauthor{Quinet}\\,(\\citeyear{Quinet}) show the best general agreement with the experimental values presented in this and previous studies. The agreement is good with the exception of the b\\,$^2$H$_{11/2}$ level (\\citeauthor{Rostohar}\\,\\citeyear{Rostohar}) which differs by $5\\sigma$. This deviation has been explained by \\citeauthor{Hartman}\\,\\citeyear{Hartman} as an effect of level mixing between the b\\,$^2$H$_{11/2}$ and the a\\,$^4$G$_{11/2}$ levels. In addition, this mixing has been observed by \\citeauthor{Johansson}\\,(\\citeyear{Johansson}) through unexpected spin forbidden spectral lines, e.g. the b\\,$^2$H$_{11/2}$ - z\\,$^6$F$_{9/2}$ transition at 6269.97 \\AA . Comparison of the theoretical lifetime values in Table~\\ref{Lifetimes} for a\\,$^4$G$_{11/2}$ and b\\,$^2$H$_{11/2}$ indicate that the lifetime of the a\\,$^4$G$_{11/2}$ level should be approximately one tenth of the lifetime of b\\,$^2$H$_{11/2}$. However, the experimental values reveal that the a\\,$^4$G$_{11/2}$ level is only one fifth of the lifetime for b\\,$^2$H$_{11/2}$ indicating that the mixing effect can be observed in the lifetimes. Calculations are sensitive to level mixing and performing {\\it ab initio} calculations which reproduce the mixing properties of the system is not trivial. This is illustrated in \\citeauthor{Nahar} (\\citeyear{Nahar}), \\citeauthor{Nahar2} (\\citeyear{Nahar2}), \\citeauthor{HibbertFe} (\\citeyear{HibbertFe}), \\citeauthor{Correge} (\\citeyear{Correge}) and \\citeauthor{Pickering_mixing} (\\citeyear{Pickering_mixing}). In particular, \\citeauthor{Pickering_mixing} (\\citeyear{Pickering_mixing}) emphasized that even though the agreement between the theoretical and experimental \\ion{Fe}{ii} $A$-values investigated \\citeauthor{Pickering_mixing} (\\citeyear{Pickering_mixing}) are in general extremely good, probabilities related to spin forbidden transitions that occur due to level mixing may differ by up to an order of magnitude." }, "0910/0910.4590_arXiv.txt": { "abstract": "The WMAP haze is an excess in microwave emission coming from the center of the Milky Way galaxy. In the case of synchrotron emission models of the haze, we present tests for the source of radiating high-energy electrons/positrons. We explore several models in the case of a pulsar population or dark matter annihilation as the source. These morphological signatures of these models are small behind the WMAP Galactic mask, but are testable and constrain the source models. We show that detailed measurements of the morphology may distinguish between the pulsar and dark matter interpretations as well as differentiate among different pulsar models and dark matter profile models individually. Specifically, we find that a zero central density Galactic pulsar population model is in tension with the observed WMAP haze. The Planck Observatory's greater sensitivity and expected smaller Galactic mask should potentially provide a robust signature of the WMAP haze as either a pulsar population or the dark matter. ", "introduction": "The Wilkinson Microwave Anisotropy Probe (WMAP) has provided a detailed map of the cosmic microwave background (CMB) as well as the foreground from our Galaxy~\\cite{Hinshaw08,Gold08}. The foreground emission from the center of the Galaxy shows an excess of emission in the inner $5-20\\degree$ around the Galactic center. This excess of microwave emission is known as the ``WMAP haze''~\\cite{Finkbeiner04}. Due to the Galactic plane, the current data on the haze only goes down to about 6 degrees from the Galactic center. Two notable features of the haze are the approximate spherical symmetry and the strong angular dependence of the flux. In particular, the flux increases quickly toward the Galactic center. The size and shape of the haze is currently not well constrained. Using a different foreground model may change the look of the haze slightly. For instance, Ref.~\\cite{Bottino08} used a different synchrotron map to model the WMAP data and found that the haze in that case is smaller in spatial extent and slightly smaller in magnitude. A cleaner foreground subtraction is needed to further constrain the source of the haze. In this paper, we will use the measurement of the haze presented in Ref.~\\cite{Hooper07}. An explanation for the WMAP haze is a previously unknown source of microwave emission. The frequency dependence of the haze is quite hard, so a hard source like synchrotron radiation is a likely candidate~\\cite{Finkbeiner07,Dobler07,Hooper07}. The magnetic field in the Galactic center tends to be tens of microgauss. Therefore, the synchrotron radiation from highly relativistic electrons and positrons near the Galactic center could be the source of the microwave signal~\\cite{Ferriere09}. Another possibility for the source of the haze was thermal bremsstrahlung from ionized gas~\\cite{Finkbeiner04}. However, the H$\\alpha$ skymap shows no significant increase in the regions where the haze is strongest. At high density and high temperature ($\\sim 10^{5}\\rm K$) such a gas could still explain the haze, but the emission from such regions is constrained to be small, ruling it out as the source of the haze~\\cite{Kurtz94}. In the synchrotron emission interpretation, calculating the complete propagation of electrons in the interstellar medium requires the full diffusion-loss equation. This includes spatial diffusion of the $e^{+}e^{-}$, reacceleration of the particles due to momentum-space diffusion, energy loss due to several different mechanisms, convection of the particles in the Galaxy, and the particle source. To solve this complete equation, one may use the software GALPROP~\\cite{Strong98,Strong00}. However, the major necessary physical features of the source and synchrotron emission can be sufficiently modeled using a Green's function solution to the diffusion-loss equation, which we employ~\\cite{Baltz04,Baltz98}. The source of high energy electrons and positrons required for synchrotron emission remains a mystery~\\cite{Zhang08}. One intriguing explanation of the haze is from dark matter annihilations in the Galactic center~\\cite{Finkbeiner07}. The $e^{+}e^{-}$ produced from by the annihilation products move through the Galactic magnetic field, creating the diffuse synchrotron emission~\\cite{Hooper07}. This model matches the haze well, especially the strong angular dependence and approximate spherical symmetry. It should be noted that Ref.~\\cite{Cumberbatch09} claims that dark matter annihilations cannot explain the haze because it would require too large of a clumpiness boost factor. In this work, we employ similar methods to those used in that paper and do not find that too large of a boost factor is necessary. Ref.~\\cite{Cumberbatch09} averaged the dark matter density over the direction perpendicular to the Galactic plane, which is inaccurate in light of the morphological effects we present below. Another source for the haze is $e^{+}e^{-}$ coming from the magnetosphere boundary of Galactic pulsars~\\cite{Kaplinghat09}. This type of emission from nearby pulsars can be the source of positrons in the PAMELA results~\\cite{Adriani08,Boulares89,Hooper08,Yuksel08,Profumo08} and also the source of features seen by ATIC~\\cite{Chang08,Yuksel08,Profumo08} and the Fermi Gamma-Ray Space Telescope~\\cite{Abdo09,Grasso09}. In this paper, we explore models of the WMAP haze in the context of synchrotron radiation from the electron and positron production of Galactic pulsars and dark matter annihilations. We will show tests of this diffuse synchrotron emission which could distinguish between the exotic explanation of dark matter annihilations and the astrophysical pulsar sources. We also show how haze observations can test pulsar and dark matter models separately. These tests can be done with upcoming results from the Planck Observatory, which has been shown, e.g.\\ in Leach, et al.~\\cite{Leach08}, to be expected to have a much smaller Galactic mask than WMAP~\\cite{Planck06,Stolyarov01}. Recently, Kaplinghat et al.~\\cite{Kaplinghat09} showed that the morphology of the haze is different depending on the source. This work pointed out that for a source with spherical symmetry, the haze should be slightly elliptically stretched along the Galactic plane while for a centrally-peaked pulsar source, the signal lies primarily along the Galactic plane and has less spherical symmetry. In this work, we examine the morphological structure of the WMAP haze in detail. First, we consider the signal due to pulsars, both from a Gaussian distribution and one which vanishes at the Galactic center. Second, we look at dark matter annihilations models' sensitivity to the dark matter density profile. Lastly, we compare the pulsar and dark matter scenarios to quantitatively distinguish them from one another. ", "conclusions": "High-energy electrons/positrons moving in the Galactic magnetic field could be the source of the WMAP haze. Using the diffusion-loss equation, we have taken a set of models of likely sources and calculated the resulting synchrotron signal as a test for the models. The simplified equation can be solved analytically using a Green's function approach, where it is useful for an understanding of the underlying physics. We considered tests of several models of the haze source, namely a peaked Gaussian distribution of pulsars, a distribution of pulsars peaked away from the Galactic center, and two dark matter distributions as possible sources, all consistent with other constraints. We find that the WMAP haze could be caused by diffuse synchrotron emission due to pulsars, in agreement with Kaplinghat et al.~\\cite{Kaplinghat09}. Moreover, we find that if the haze is caused by pulsars, the angular flux profile from the Galactic center should be peaked more sharply when using a line-of-sight away from the Galactic plane and is flattened significantly when using a line-of-sight closer to the Galactic plane. Also, if the pulsar distribution vanishes in the Galactic center, a brightness about $20\\degree$ out along the Galactic plane should be visible. With annihilating dark matter, the angular flux profile from Galactic center is largely independent of direction and has rough spherical symmetry. If the signal is found to be strongly spherically symmetric, then this would be an indication that annihilating dark matter could be the true source. This directional-dependence of the angular flux profile would be a smoking gun for pulsars causing the haze as opposed to the more spherical dark matter explanation. The Planck probe's increased sensitivity, larger number of bands, enhanced models, and expected smaller Galactic mask will test these models~\\cite{Leach08,Planck06,Stolyarov01}. This will be instrumental in determining the source of the WMAP haze as an astrophysical signal or the indirect detection of the dark matter." }, "0910/0910.4559_arXiv.txt": { "abstract": "The geometry of a light wavefront evolving in the 3--space associated with a post-Newtonian relativistic spacetime from a flat wavefront is studied numerically by means of the ray tracing method. For a discretization of the bidimensional wavefront the surface fitting technique is used to determine the curvature of this surface at each vertex of the mesh. The relationship between the curvature of a wavefront and the change of the arrival time at different points on the Earth is also numerically discussed. ", "introduction": "\\label{intro} The description of the propagation of light in a gravitational field is even today a central problem in the general theory of relativity. The deflection of light rays and time delays of electromagnetic signals due to the presence of a gravitational field are phenomena detectable with current experimental techniques which allow design new tests for general relativity. In this line, Samuel~\\cite{Sam} recently proposed a method for the direct measurement of the curvature of a light wavefront initially flat, curved when light crosses regions where the gravitational field is non vanishing. He found a relationship between the differences of arrival time recorded at four points on the Earth, measured by employing techniques of very long base interferometry and the volume of a parallelepided determined by four points in the curved wavefront surface. This surface is described by means of a polynomial approximation of the eikonal in a Schwarzschild gravitational field. For more complex gravitational models, such as those considered by Klioner and Peip~\\cite{KP}, de Felice {\\it et al.}~\\cite{dF} or Kopeikin and Sch\\\"afer~\\cite{KS} in studies of light propagation in the solar system, the use of numerical methods for the determination of the geometry of the wavefront surface would also be required. An analytical approach to the relativistic modeling of light propagation has also been developed recently by Le Poncin-Lafitte {\\it et al.}~\\cite{Lep} and Teyssandier and Le Poncin-Lafitte~\\cite{Tey}, where they present methods based on Synge's world function and the perturbative series of powers of the Newtonian gravitational constant, to determine the post-Minkowskian expansions of the time transfer functions. Nowadays there are numerous techniques in computational differential geometry which allow to analyze geometric properties of surfaces embedded in the ordinary Euclidean space. Techniques of this type are widely applied in different areas such as Computational Geometry, Computer Vision or Seismology. In one of these methods, developed in works by Garimella and Swartz~\\cite{GS} and Cazals and Pouget~\\cite{CP}, the estimation of differential quantities is established using a fitting of the local representation of the surface by means of a height function given by a Taylor polynomial. A survey of methods for the extraction of quadric surfaces from triangular meshes is found in Petitjean~ \\cite{Pet}. In this work we consider a discretization of the wavefront surface, replacing this surface by a polyhedral whose faces are equilateral triangles. At initial time, the surface is assumed to be flat and far enough from a gravitational source (say, the Sun) and moving towards this source. We study the deformation of the instantaneous polyhedral representing the wavefront when crossing a region in the relativistic 3--space near the Sun due to the bending of light rays by the gravitational field. In this study, we apply the ray tracing method with initial values on the vertices of the triangular mesh to obtain the corresponding discrete surface at each instant of time. Then we apply the techniques given in \\cite{GS} and \\cite{CP} to describe the wavefront as a surface embedded in the Riemannian 3--space of the post-Newtonian formalism of general relativity. For each vertex in the instantaneous mesh we obtain a quadric which represents locally the surface by applying the least-squares method to the immediate neighboring vertex around the considered point which is represented in normal coordinates adapted to the light rays. The structure of the paper is as follows: In Section 2 we briefly introduce the basic model for the wavefront propagation in the post-Newtonian formalism. In Section 3, we establish a discretized model of the wavefront surface by means of a regular triangulation and we describe the method employed in this work for the study of the curvature of this surface. In Section 4, a numerical estimation of the curvature of the surface is derived using the ray tracing method. For the numerical integration of the light ray equation, we use the {\\tt Taylor} algorithm implemented by Jorba and Zou \\cite{JZ} which is based on the Taylor series method for the integration of ordinary differential equations and which allow the use of high order numerical integrators and arbitrary arithmetic accuracy, as is required to describe the influence of weak gravitational perturbations on the bending of light rays. Finally, in Section 5, we apply the method discussed above to study the effect of the wavefront curvature on the variation of the arrival time of the light at points on the Earth surface, following the model proposed in Samuel's test \\cite{Sam}. The paper concludes by giving another approach to the estimation of the curvature of the wavefront, derived from an approximation of the Wald curvature \\cite{Blu} associated with a quadruple of points in the wavefront. ", "conclusions": "The ray tracing numerical method provides a useful tool for the description of spacelike bidimensional wavefronts within the framework of the general relativity. We have studied a method, based on techniques of computational geometry, that allows to estimate the curvature properties of the surface by making a least-squares fitting of the wavefront surface by a quadric surface in the neighborhood of each point of this surface. The computation of the light rays is carried out using an algorithm based on the Taylor method for the solution of differential equations and employing high arithmetic precision. Further, we have applied a projection at each step of the numerical integration process that allows to guarantee the fulfillment (at machine precision) of the isotropy condition for the tangent vector to the light ray. We have also studied numerically the dependence of the curvature properties of the wavefront surface on the value of the Eddington parameter $\\gamma$. On the other hand, we have employed a geometric computational approach to the study of the model proposed by Samuel as a new general relativity test, by determining a numerical approximation of the volume corresponding to a tetrahedron formed by four points on the wavefront that reaches four receiving stations on the Earth surface. Finally, we have obtained an estimation of the Wald curvature for the wavefront in a vicinity of the Earth by using the differences of arrival time recorded at four receiving stations on the Earth." }, "0910/0910.3465_arXiv.txt": { "abstract": "We have performed an On-The-Fly (OTF) mapping survey of ${\\rm ^{12}{CO(J=1-0)}}$ emission in 28 Virgo cluster spiral galaxies using the Five College Radio Astronomy Observatory (FCRAO) 14-m telescope. This survey aims to characterize the CO distribution, kinematics, and luminosity of a large sample of galaxies covering the full extents of stellar disks, rather than sampling only the inner disks or the major axis as was done by many previous single dish and interferometric CO surveys. CO emission is detected in 20 galaxies among the 28 Virgo spirals observed. An atlas consisting of global measures, radial measures, and maps, is presented for each detected galaxy. A note summarizing the CO data is also presented along with relevant information from the literature. The CO properties derived from our OTF observations are presented and compared with the results from the FCRAO Extragalactic CO Survey by Young et al. (1995) which utilized position-switching observations along the major axis and a model fitting method. We find that our OTF derived CO properties agree well with the Young et al. results in many cases, but the Young et al. measurements are larger by a factor of 1.4 - 2.4 for seven (out of 18) cases. We will explore further the possible causes for the discrepancy in the analysis paper currently under preparation. ", "introduction": "The high galaxy density and the proximity make the Virgo cluster a particularly interesting laboratory for a galaxy evolution study. Its dynamical evolution is still in progress, and evidence for significant environmental effects is ubiquitous \\citep[e.g., ][]{chung07}. The nearness of the Virgo cluster makes it possible to observe its member galaxies with excellent spatial resolution ($1\\arcsec\\ \\sim 90$ pc). It is the first cluster for which significant HI imaging was done \\citep{vgo84}, and various new HI surveys such as ALFALFA \\citep{gio07} and VIVA \\citep{chu07} have recently been conducted. The Virgo cluster has also been the subject of many recent multi-wavelength surveys, such as in radio continuum by NRAO VLA Sky Survey \\citep[NVSS;][]{con98}, in H$\\alpha$ \\citep{koo04,che06}, in UV by FAUST \\citep{bro97} and Galaxy Evolution Explorer (GALEX) \\citep[e.g.,][]{dal07}, and in IR by Spitzer \\citep{ken08}. High quality optical images are also available from the Hubble Space Telescope (HST)/ACS Virgo Cluster Survey \\citep{cot04} and Sloan Digital Sky Survey (SDSS). Here, we present the results from a new imaging survey of $J=1\\rightarrow0\\ ^{12}$CO emission in a large sample of Virgo galaxies in order to address the distribution and characteristics of dense molecular gas in these galaxies. It is well established that molecular clouds are the sites of ongoing star formation (\\citet{lar03} and references therein), and carbon monoxide (CO) is the most commonly used tracer of molecular hydrogen (H$_2$), which is the most abundant but invisible component of cold and dense clouds \\citep[e.g.,][]{sol91}. The global H$_2$ content and its distribution in galaxies and a comparison with other gas and stellar components as a function of morphological type, luminosity, and environment are some of the key insights one can derive from CO observations (see \\citet{you91} and references therein). Existing CO surveys of Virgo cluster galaxies suffer from limited spatial coverage and small sample sizes. The FCRAO extragalactic survey \\citep{young95} is the first CO survey covering a wide range of distance, diameter, morphological type, and blue luminosity of some 300 galaxies conducted using the FCRAO 14-m telescope, including CO measurements (detections and upper limits) of 65 Virgo galaxies. Because of the large observing time required, these observations were conducted in the position-switching mode, primarily along the optical major axis of the disks, and the global CO line luminosity was derived assuming a model distribution. More recently, high angular resolution CO images have been obtained using interferometric measurements by \\citet{sak99} and \\citet{sof03}. These measurements reveal a detailed molecular gas distribution at 100 pc scales, but they are limited only to the central region of galaxies. The BIMA SONG (Survey Of Nearby Galaxies; \\citet{hel03}) has also carried out an imaging survey of CO emission in several Virgo spiral galaxies, incorporating the short spacing data from the NRAO 12-m telescope. The total number of Virgo galaxies imaged by the BIMA SONG is small (six), however. These earlier CO observations have revealed important insights on the molecular interstellar medium (ISM) in these galaxies. For example, CO emission is often centrally concentrated, in contrast to the centrally deficient HI distributions commonly found in these galaxies. The molecular gas distribution also shows little evidence for any influence of the gas stripping mechanisms \\citep[e.g.,][]{ken89}. No central CO peak \\citep{hel03} or nuclear molecular rings \\citep[e.g.,][]{ion05} are seen in other cases. The influence of cluster environment on the molecular content is still poorly understood \\citep[e.g.,][]{bos06}. A distinguishing characteristic of our new CO survey is the complete imaging of $^{\\rm 12}$CO (J=1--0) emission of a large sample (28 galaxies) of Virgo spirals using the On-The-Fly (OTF) mapping mode of the Five College Radio Astronomy Observatory (FCRAO) 14-m telescope. Our map size of $10^{\\prime} \\times 10^{\\prime}$ is larger than the optical diameter $\\rm D_{25}$ and it covers the entire stellar disk of each galaxy. The 45\\arcsec\\ angular resolution of the new CO images is well matched to the existing VLA HI data and is well suited for comparison with other high resolution multi-wavelength data. The specific questions we aim to address are : 1) To what extent is the CO distribution governed by the disk dynamics? 2) Is there a clear phase transition between HI and $\\rm H_{2}$ as a function of the interstellar radiation field? 3) Are there any systematic differences in the CO properties of spiral galaxies in different environments? We present in this paper the data and the CO atlas from our OTF mapping survey. In section 2, we describe the sample selection. Observations and data reduction process are described in Section 3. The CO atlas and CO properties are presented in Section 4, and our results are compared with those of \\citet{young95}. Molecular gas distribution in individual galaxies is discussed in Section 5, and the summary and conclusion are given in Section 6. The analysis and interpretation of the data addressing the above questions will be presented in our subsequent papers. ", "conclusions": "" }, "0910/0910.3186_arXiv.txt": { "abstract": "We have analysed the XMM-Newton spectra of SS\\,433 using a standard model of adiabatically and radiatively cooling X-ray jets. The multi-temperature thermal jet model reproduces well the strongest observed emission line fluxes. Fitting the He- and H-like iron line fluxes, we find that the visible blue jet base temperature is $\\approx 17$~keV, the jet kinetic luminosity $L_k \\sim 2 \\cdot 10^{39}$\\,erg/s and the absorbing column density $N_H \\sim 1.5 \\cdot 10^{22}$\\,cm$^{-2}$. All these parameters are in line with the previous studies. The thermal model alone can not reproduce the continuum radiation in the XMM spectral range, the fluorescent iron line and some broad spectral features. Using the thermal jet-plus-reflection model, we find a notable contribution of ionized reflection to the spectrum in the energy range from $\\sim 3$ to 12~keV. The reflecting surface is highly ionized ($\\xi \\sim 300$), the illuminating radiation photon index changes from the flat spectrum ($\\Gamma \\approx 2$) in the 7\\,--\\,12~keV range to $\\Gamma \\approx 1.6$ in the range of 4\\,--\\,7~keV, and to $\\Gamma \\la 1$ in the range of 2\\,--\\,4~keV. We conclude that the reflected spectrum is an evidence of the supercritical disc funnel, where the illuminating radiation comes from deeper funnel regions, to be further reflected in the outer visible funnel walls ($r\\ga 2 \\cdot 10^{11}$ cm). In the multiple scatterings in the funnel, the harder radiation $> 7$~keV may survive absorption, but softer radiation is absorbed, making the illuminating spectrum curved. We have not found any evidences of reflection in the soft 0.8\\,--\\,2~keV energy range, instead, a soft excess is detected, that does not depend on the thermal jet model details. However the soft component spectrum is basically unknown. This soft component might prove to be the direct radiation of the visible funnel wall. It is represented here either as black body radiation with the temperature of $\\theta_{bb} \\approx 0.1$~keV and luminosity of $L_{bb} \\sim 3 \\cdot 10^{37}$\\,erg/s, or with a multicolour funnel (MCF) model. The soft spectral component has about the same parameters as those found in ULXs. ", "introduction": "\\label{S:Intro} SS\\,433 is the only known persistent superaccretor in the Galaxy -- a source of relativistic jets \\citep[][for review]{Fab04}. This is a massive close binary, where the compact star is most probably a black hole \\citep{Giesetal02,Cheretal05,Hill_Gies08,Blundelletal08}. SS\\,433 intrinsic luminosity is estimated to be $\\sim 10^{40}$\\,erg/s, with its maximum located in non-observed UV region. In effect this is a very blue, but heavily absorbed object; its estimated temperature depends on the accretion disc orientation \\citep{Murdin_Clark80,Cheretal82,Dolanetal97}, being $\\sim 50000 - 70000$\\,K when the disc is the most open to the observer. It is important that almost all the observed radiation is formed in the supercritical accretion disc, and the donor star contributes less than 20\\,\\% of the optical radiation. The system's extreme luminosity suggests that the black hole's mass is $\\sim 10$\\,M$_{\\sun}$. Such a big energy budget of the object is supported by a very well measured kinetic luminosity of the jets, $\\sim 10^{39}$\\,erg/s, both in direct X-ray and optical studies of the jets and in the studies of the jet-powered nebula W\\,50 \\citep{Fab04}. At the same time we know that practically all the energy at the accretion onto a relativistic star is released in X-rays. This means that the observed radiation of SS\\,433 was thermalised in the strong wind coming from the supercritical disc. The wind from a supercritical disc was first predicted by \\citet{ShakSun73} and later confirmed in radiation-hydrodynamic simulations \\citep{Eggumetal88,Okudaetal05,Ohsuga05}. These ideas led to a prediction that SS\\,433, being observed face-on appears as an extremely bright X-ray source, and we may expect an appearance of a new type of X-ray sources in galaxies \\citep{Katz87,FabMesch00,FabMesch01} -- the face-on SS\\,433 stars. It is quite possible that the new type of X-ray sources, ULXs (ultraluminous X-ray sources, \\citet{Roberts07}) are SS\\,433-like objects observed nearly face-on. Their observed X-ray luminosities are $10^{39} - 3 \\cdot 10^{41}$\\,erg/s and they are certainly related to the massive star population. The disc orientation in SS\\,433 is about edge-on, the angle between the disc plane and the line of sight is $10 \\pm 20^{\\circ}$, due to the precessional variations. Therefore we have no chance to observe the funnel in the supercritical disc directly. The supercritical disc radiation is not isotropic. If one takes into account the geometrical beaming in the disc funnel with the (half) opening angle $\\vartheta_f \\sim 25^{\\circ} - 30^{\\circ}$ \\citep{Ohsuga05} and that a supercritical disc surface radiation is local Eddington \\citep{ShakSun73}, one finds \\citep{Fab_etal06}, that the observed X-ray luminosity of $\\sim 10^{41}$\\,erg/s is expected for the face-on SS433 star. The recent data show \\citep{Stobbart06,Berghea08} that ULXs posses curvature and rather flat X-ray spectra, which are difficult to interpret with a single-component or any other simple model. The supercritical accretion discs are expected to have flat ($\\nu F_{\\nu} \\propto \\nu^0$) X-ray spectrum \\citep{Poutan07}, because the energy release $Q(r)$ in the discs must be $Q(r) \\propto r^{-2}$ ($T(r) \\propto r^{-1/2}$) to make the disc thick due to the radiation pressure. The X-ray luminosity of SS\\,433 is $\\sim 10^{36}$\\,erg/s (at a distance of 5.5\\,kpc \\citep{Blund_Bowler04}), four orders of magnitude less than the bolometric luminosity. It is believed that all the X-rays come from the cooling X-ray jets. The jet may be easily accelerated by the radiation in the hydrodynamic funnel \\citep[e.g., ][]{Ohsuga05} of the supercritical disc. The jet velocity value, $v_j \\approx 0.26 c$, and its unique stability, where the velocity does not depend on the activity state, indicate that the jet acceleration must be controlled by the line-locking mechanism \\citep{shapiro86,Fab04}. The funnel has to be relatively transparent for the accelerating radiation, as it was confirmed in radiation-hydrodynamic simulatuons \\citep[e.g., ][]{Ohsuga05}. These reasonings, however, bear a problem, why do not we observe any funnel radiation, missing four orders of magnitude in the X-rays? Even a small amount of gas in the most outer part of the funnel, where we observe the X-ray jet base, may reflect and scatter some part of the direct $\\sim 10^{40}$\\,erg/s of the funnel radiation. The structure of the funnel and the jet acceleration/collimation region is unknown. Apparently, the direct funnel radiation is entirely blocked for the observer. The 'standard' jet model \\citep{brinkmann88,kotani96,marshall02,filippova06} suggests that all the observed X-ray radiation of SS\\,433 is formed in the cooling X-ray jets. However, an analysis of the latest XMM observations \\citep{brinkmann05} led the authors to a conclusion that the 'standard' jet model can not produce the observed X-ray continuum. In the 'standard' jet model there are two conical anti-parallel jets, considered identical. The jet's gas is optically thin, being in collisional ionization equilibrium and cooling adiabatically (or adiabatically plus radiatively). The jets are observed beginning from the jet base $r_0$ (the distance between the base of the visible jet and the black hole, different for the red and blue jets), which depends on the precessional phase and on the eclipses by the donor star. The gas temperature at the jet base $T_0(r_0)$ is estimated from observations. Precessional and orbital variability of SS\\,433 is well-known \\citep{Fab04}, the jets (and both the accretion disc, and the accretion disc wind) precess with a 164-day period and an amplitude of $\\pm 20^{\\circ}$. At the precessional phase $\\psi \\approx 0$ the disc is the most open to the observer, the angle between the jets and line of sight is $\\sim 60^{\\circ}$ (we refer here to the approximate values because there are nutation-like and sporadic variabilities, both of about $3-5^{\\circ}$). The accretion disc rim (or the opaque part of the wind) is thick $h/r \\sim1$ \\citep{filippova06}. Comparing the amplitudes of the precessional and orbital variabilities in different X-ray bands, \\citet{Cheretal05} found that both the donor star and the outer disc rim have about the same size, Therefore, one may expect the same radius for the approaching jet base $r_0 \\sim 10^{12}$\\,cm. In X-ray observations with ASCA and Chandra \\citep{kotani96,marshall02}, dozens of X-ray emission lines formed both in blue and red jets were resolved. The lines are variable both in positions (in accordance with the jet kinematic model) and in intensities, the strongest are He- and H-like $\\text{Fe XXV\\,K}\\alpha$ and $\\text{Fe\\,XXVI\\,L}\\alpha$ iron lines. This allowed the 'iron line diagnostics' of the gas temperature at the jet bases by measuring the line ratio Fe\\,XXV/Fe\\,XXVI. The temperature at the base of the jet was estimated as $\\theta_0 = kT_0 = 10 \\div20$~keV \\citep{kotani96,marshall02,namiki03}. In the latest XMM-Newton observations \\citet{brinkmann05} estimate the temperature of about $\\theta_0 \\sim 17 \\pm 2$~keV, with the main uncertainty coming from the form of the underlying continuum. Fitting the continuum with a thermal bremsstrahlung model \\citep[GINGA data, ][]{brinknann91} gives a notably bigger value, $\\theta_0 \\ga 30 $~keV. The jet base temperature determined both in the line diagnostics and in the continuum fits, drops notably in eclipses, what confirms the cooling jet model. Analysing the numerous ASCA data, \\citet{kotani96} found that the farther part of the receding jet is weakened (probably absorbed in the external gas located in the disc plane) and Nickel is highly overabundant ($\\sim 10$ times) in the jets. The last finding has been confirmed in the XMM observations \\citep{brinkmann05}. Note that some discrepancies may exist as the system SS\\,433 is highly variable and its appearance depends strongly on the precessional phase. \\citet{brinkmann05} found that the XMM SS\\,433 spectrum can not be fitted with any simple continuum law. The numerous lines point at the thermal model, but a strong continuum curvature indicates the Comptonization. The best formal model of the additional continuum component is a broken power law with a break at $\\sim 7.1$~keV, where the low energy PL rises with energy ($\\Gamma \\sim -1$), the high energy part is rather steep ($\\Gamma \\sim 4$), however, the parameters are not usually well determined in the fits. The overall continuum is too hard to be bremsstrahlung, the electron-electron bremsstrahlung can not fit it even for the temperatures of 40~keV. A highly absorbed additional thermal component gives a too strong $\\sim 7$~kev edge. \\citet{brinkmann05} discuss a model with a single, very broad $\\sim 7$~kev line, which may be formed in the innermost hot jet (otherwise hidden) and Compton down-scattered by colder material. These uncertainties in continuum make problems for the iron-line diagnostics of the jet base temperature. Practically in all the studies of SS\\,433 X-ray spectra the jet mass loss $\\dot M_j$ and the jet kinetic luminosity $L_k = \\dot M_j v_j^2 /2$ were determined. To estimate $L_k$ one needs to adopt the jet (half) opening angle $\\vartheta_j$. The mass loss rate derived from the X-ray line intensities depends strongly on the opening angle adopted, i.~e. gas density at the jet base, because the emission measure is $\\propto n^2$. The kinetic luminosity values derived from X-ray spectra were quite diverse, up to $10^{42}$\\,erg/s. However, in two latest studies, one from the Chandra data \\citep{marshall02} and the other from the XMM data \\citep{brinkmann05}, the values are $\\sim 3 \\cdot 10^{38}$ and $\\sim 5 \\cdot 10^{39}$\\,erg/s respectively. They are even closer to one another, matching different distances to SS\\,433 adopted by these authors. \\citet{panferov} found $L_k \\sim 10^{39}$\\,erg/s from the jet Balmer line intensities in optical spectra. The mass loss rate in the jets was estimated \\citep{zealey,FabrBor87,dubner} on the base of W\\,50 nebula powered by the jets with approximately the same result, $L_k \\sim 10^{39}$\\,erg/s. Thus the jet kinetic luminosity was measured with relatively good accuracy. The jet opening angle was directly measured in the studies of X-ray and optical line widths \\citep{Fab04}. \\citet{marshall02} found $\\vartheta_j = 0.61 \\pm 0.03^{\\circ}$ from the Chandra spectra. They assumed that the density is uniform through the jet cone's cross section and $\\vartheta_j$ is the jet's cross sectional radius. \\citet{borfab87} found $\\vartheta_j = 0.82 \\pm 0.14^{\\circ}$ from optical spectra, assuming that the emitting gas density is Gaussian-distributed in the cone's cross section. We give here the values of $\\vartheta_j$ recalculated, using the \\citet{marshall02} definition. Such a coincidence in the X-ray and optical jet opening angles makes the jets pure ballistic from X-ray ($\\sim 10^{12}$\\,cm) to optical ($\\sim 10^{15}$\\,cm) regions. In optical spectra the jet opening angle has been estimated from H$\\alpha$ line profiles during 8-day long observations, the nutation broading has been taken into account and possible jet sporadic activity was diminished. In other Chandra observations \\citep{namiki03,lopez06} the opening angle has been found twice as big and it was greater in the higher temperature (Fe), than in the lower temperature (Si) parts of the jets. Regarding comparatively short time scales of the X-ray observations and probable line broading due to electron scattering \\citep{namiki03} in the hottest parts of the jet, we may adopt $\\vartheta_j \\approx 0.7^{\\circ}$ as a good estimate of the jet opening angle. In this paper we analyse X-ray spectra of SS\\,433 using XMM data and the 'standard' jet model. We do not try to determine the jet kinetic luminosity, because the X-ray continuum is complex and can not be explained in the 'standard' jet model \\citep{brinkmann05}. Instead we adopt these two main parameters -- $L_k = 10^{39}$\\,erg/s ($\\dot M_j = 3.3 \\cdot 10^{19}$\\,g/s) and $2\\vartheta_j = 1.5^{\\circ}$, as well established and known. We find that the line fluxes can be well matched with the 'standard' jet model and study the additional components of the X-ray spectrum of SS\\,433. ", "conclusions": "The goal of the present paper was to find indications of the supercritical accretion disc funnel in SS\\,433 X-ray radiation. The bolometric luminosity of the disc in SS\\,433 is $\\sim 10^{40}$\\,erg/s and all the luminosity must be initially released in X-rays. This is confirmed by the well-established kinetic luminosity of SS\\,433 jets, which is $\\sim 10^{39}$\\,erg/s, and the fact that the jets are formed in the deepest places of the disc funnel close to the black hole. The observed X-ray luminosity of SS\\,433 is $\\sim 10^{36}$\\,erg/s and it is the radiation of the cooling X-ray jets. Both the orientation of SS\\,433 and the visibility conditions do not allow us to see any deep areas of the funnel. However, the strong mass loss of the accretion disc gives us a hope for detecting some indications of the funnel radiation. We have analysed the XXM spectra of SS\\,433 with a well-known standard model of the adiabatically cooling X-ray jets, taking into account cooling by radiation. We confirm that the thermal jet model reproduces the emission line fluxes quite well. When the disc is most open to the observer, the visible blue jet base is $r_{0} \\approx 2\\cdot10^{11}$\\,cm, the gas temperature at $r_{0}$ is $\\theta_0 \\approx 17$~keV. The IS gas column density is $N_H \\sim 1.5 \\cdot 10^{22}$\\,cm$^{-2}$, which is in good agreement with the IS extinction found in optical and UV observations. We confirm the previous finding \\citep{kotani96,brinkmann05} that the red jet is probably not seen (blocked) in the soft energy range of 0.8~--2.0~keV. However, both the model with blocked the red jet portions, and the model with the whole red jet visible (but with some higher $N_H$) give the same result, that the thermal jet model alone can not explain the soft continuum. We also confirm that Nickel is highly overabundant in the jets \\citep{kotani96,brinkmann05}. We have adopted the luminosity $L_k = 10^{39}$\\,erg/s and the jet opening angle $2\\vartheta_j = 1.5^{\\circ}$ as well established and known. However, the derived jet visible base $r_{0}$ is too small, it has to be $\\sim 10^{12}$\\,cm to satisfy the observed eclipses by the donor star. Using the scaling formula $r_0 \\propto \\dot M_j^2 / \\Omega_j$ one may find that during these XMM observations the jet kinetic luminosity was $L_k \\sim 2 \\cdot 10^{39}$\\,erg/s. This scaling should not change the thermal jet spectrum. We find that the thermal jet model alone can not reproduce the continuum radiation in the XMM spectral range. Therefore we use the thermal jet model together with the REFLION ionized reflection model and find quite a good representation of the spectra. Introducing the reflection not only explains the continuum curvature at the energies of $> 3-4$~kev and the fluorescent line of quasi-neutral iron, it reproduces the broad absorption feature at $\\sim 8$~keV recently detected by \\citet{kubota07}. We independently study three energy ranges (2\\,--\\,4, 4\\,--\\,7 and 7\\,--\\,12~keV) with the same thermal jet model to find the reflection parameters. We find that the ionization parameter (\\ref{E:ionpar}) is about the same in these ranges, $\\xi \\sim 300$, which indicates a highly ionized reflection surface. The illuminating radiation photon index changes from flat, $\\Gamma \\approx 2$, in the 7\\,--\\,12~keV range to $\\Gamma \\approx 1.6$ in the range 4\\,--\\,7~keV and to $\\Gamma \\la 1$ in the range 2\\,--\\,4~keV. We conclude that the additional reflected spectrum is an indication of the funnel radiation. The illuminating radiation spectrum is flat in the range of 7\\,--\\,12~keV. With multiple scatterings in the funnel the hard radiation may survive absorption. The reflected luminosity in this 7\\,--\\,12~keV spectral range is $L_{refl} \\sim 10^{36}$\\,erg/s. The softer ($2-7$~keV) part of the illuminating spectrum (Fig.\\,\\ref{F:add}) carries a trace of absorption. This is assumed to be due to the multiple scatterings in the funnel. It is important to note that an existence of He- and H-like absorption edges at about zero velocity is a mandatory property of the funnel spectrum to be able to produce the observed jet velocity due to the line-locking mechanism. The supercritical accretion disc spectrum is expected to be flat, however, it extends up to a few keV only. Comptonization may extend and flatten the spectrum to higher energies. The interaction of the high velocity gas moving along the jet axis with the walls' wind makes the conditions for the Comptonization throughout the whole funnel length consistent. Direct evidences of the Comptonized radiation have recently been found in the INTEGRAL observations \\citep{Cheretal05}. We have not found any evidences of the reflection in the soft 0.8\\,--\\,2.0~keV energy range. The soft excess is detected in the SS\\,433 spectra. We suppose that in the soft X-rays we observe direct radiation of the visible funnel wall. We represented this component as a black body radiation with a temperature of $\\theta_{bb} \\approx 0.1$~keV and luminosity of $L_{BB} \\sim 3 \\cdot 10^{37}$\\,erg/s. The soft excess may be also fitted with a multicolour funnel (MCF) model. Both the soft X-ray luminosity and its derived spectrum strongly depend on the column density $N_H$. The soft component is necessary to add in order to understand the SS\\,433 spectra within the framework of the thermal jet model. The soft excess is observed in the spectra of ULXs with the soft component temperature of $\\theta \\sim 0.1$~keV, which is identical with that found in SS\\,433. If the ULXs or some of them are nearly face-on versions of SS\\,433, one may adjust their soft X-ray components to the outer funnel walls radiation \\citep{Poutan07}." }, "0910/0910.1183_arXiv.txt": { "abstract": "The Exo-Planet Imaging Camera and Spectrograph (EPICS) for the future 42-meter European-Extremely Large Telescope, will enable direct images, and spectra for both young and old Jupiter-mass planets in the infrared. To achieve the required contrast, several coronagraphic concepts -- to remove starlight -- are under investigation: conventional pupil apodization (CPA), apodized-pupil Lyot coronagraph (APLC), dual-zone coronagraph (DZC), four-quadrants phase mask (FQPM), multi-stages FQPM, annular groove phase mask (AGPM), high order optical vortex (OVC), and band-limited coronagraph (BLC). Recent experiment demonstrated the interest of an halftone-dot process -- namely microdots technique -- to generate the adequate transmission profile of pupil apodizers for CPA, APLC, and DZC concepts. Here, we examine the use of this technique to produce band-limited focal plane masks, and present guidelines for the design. Additionally, we present the first near-IR laboratory results with BLCs that confirm the microdots approach as a suitable technique for ground-based observations. ", "introduction": "By the end of 2018, challenging projects such as EPICS (Exo-Planet Imaging Camera and Spectrograph, \\citet{EPICS}) for the future 42-meter European-Extremely Large Telescope, or PFI \\citep[Planet Formation Imager,][]{M06} for the Thirty Meter Telescope (TMT) will enable direct images, and spectra for warm self-luminous and reflected-light Jovian planets. These instruments will operate with eXtreme Adaptive Optics system (XAO) designed for high Strehl ratios (i.e. $\\sim90\\%$ in $H$-band), as for SPHERE \\citep{2008SPIE.7014E..41B} and GPI \\citep{2006SPIE.6272E..18M}, forthcoming planet-finder instruments with first light planned in 2011. Several coronagraph concepts have been studied extensively in the past years, with the objective of finding optimized designs that can sufficiently suppress the on-axis starlight, allowing faint companions direct detection (e.g. \\citet{Malbet96, SIVA01, Guyon07, Corono}). Among them, the band-limited coronagraph \\citep[][BLC]{2002ApJ...570..900K} has been proposed to completely remove starlight. The BLC has the advantage of being less sensitive to the primary mirror segmentation, unavoidable with ELTs, than for other concepts \\citep{SIVA05}, and to provide achromatic behavior. Additionally, in \\citet{Corono}, we pointed out the interest of such concept on an ELT for either very bright object detection, or for the search of planets at large angular separations. Several BLC prototypes have been developed during the past years \\citep[e.g.][]{Debes04b, Trauger, Creep06, Trauger2, Moody08} for visible wavelength application. Several technical approaches have been used: (1/) gray-scale pattern written with an high-energy beam sensitive glass (HEBS) using e-beam lithography, (2/) Notch filter pattern -- binary mask -- written with thick Chromium layer on a substrate, dry-etched with high density decoupled plasma, (3/) gray-scale pattern manufactured with vacuum deposited metals and dielectrics. Even if electron-sensitized HEBS glass can accurately accommodate continuous range of transmission, the darkening of the HEBS glass under electron bombardment is accompanied by a determined phase shift, while the technique suffers from a lack of experience in the near-IR. Same constraints apply for the vacuum deposited metal technique. The notch filter has the advantage of being intrinsically achromatic. These designs, consisting of a particular implementation of small structures (stripes of opaque material with width of about some microns) must be finely controlled in size, spacing, and opacity. Mask errors and tolerance are discussed in \\citet{Kuchner03}, where requirements might be strong for near-IR application. In this paper, we examine the use of a halftone-dot process, namely microdot technique, to reproduce a continuous mask profile as already done for pupil apodizer for SPHERE \\citep{microdots1, microdots2}, and being manufactured for the JWST NIRCam coronagraph \\citep{Krist09}. These masks consist of distributions of opaque square pixels (called dots) to reproduce the continuous transmission of a filter with several advantages: relative ease of manufacture, achromaticity, reproducibility, and ability to generate continuous transmission ranges, without introducing wavefront errors. Besides, mask errors can be easily pre-compensated \\citep{2007JOSAB..24.1268D}. Section 2 describes the microdot design principle and properties, while Sect. 3 provides guidelines for the design. Section 4 presents monochromatic, and polychromatic results obtained in laboratory in the near-infrared. Finally Sect. 5 concludes on the suitability of the microdots approach for producing BLCs in the context of ground-based instruments. ", "conclusions": "We have described the development and laboratory experiment of Band-limited coronagraphs using a microdots design in the near-infrared. In this paper, we provide design guidelines, and demonstrate the microdots technique as a promising solution for BLC for ground-based observations. We have shown with numerical simulations that although total starlight cancellation is not possible, theoretical contrast offer with the microdot approach are deep enough to guarantee that the BLC will not set a limit on the performance of a ground-based instrument. We note that the theoretical treatment presented in this study does not consider the complexity introduced by potential spectral, and polarization effects of the physical mask. We identified two sampling configurations suitable for near-IR experiment: $s=16$ and $s=8$, yielding to identical performance in the experiment. Additionally, we pointed out that the interest of the microdots technique in the light of contrast and IWA requirements is not dependent of the function bandwidth ($\\epsilon$) assuming IWA $\\geq$ 3$\\lambda/D$. This already meets EPICS (20-30 mas in H-band), or SPHERE (0.1 arcsec in H-band) standard requirements. However, we note that stronger requirements of the IWA (1 or 2 $\\lambda/D$, ultimate goal of SPHERE), will set a limit on the interest of the technique. With prototypes we have demonstrated impressive performance, where limitations have been presumably (on the basis of simulations) introduced either by mask error, or wavefront error of the bench. Improvement of the results presented and resumed in Table \\ref{resum} are foreseen -- at least for the peak attenuation -- with a new set of prototypes that will provide a better accuracy of the profile in the outer part of the function (calibration issue of the manufacturing process). Additionally, these raw polychromatic results presented belong as the first tests of BLC in the near-IR, and were managed without active wavefront correction, nor particular data reduction post-processing. Performance reached in the experiment went beyond to the SPHERE prototype performances, APLC results obtained so far with a microdot apodizer on the same bench, and yield to similar contrast than the ones presented in \\citet{Creep06} (using $4^{th}$ order notch-filter mask, in a monochromatic visible-wavelength domain) although IWA are not exactly identical. Ultimately, these final prototypes will be implemented on HOT \\citep[the High-Order Testbench, the AO-facility at ESO,][]{HOTbench}, and being compared with others \\citep[FQPM, APLC, and Lyots,][]{HOTcorono, microdots1} with atmospheric turbulence generator and AO-correction, as already initiated with intensive simulations \\citep{Corono}. Results of this experiment will be presented in a forthcoming paper. Although this study was carried out in the context of R$\\&$D activities for EPICS, it is potentially applicable to upcoming instruments such as SPHERE, or GPI. We note that the interest of the technique presented in the paper for space-based operations, is subject to science cases (IWA and contrast requirements). For instance, high star-planet contrast ($10^{-10}$ in the visible) at less than 0.1$\\arcsec$ (Terrestrial planet), will very likely required a wavefront stability only reachable with a space telescope. In that situation, the technique employed here will impose a limit on the accessible contrast. Notch-filter masks might be more appropriate." }, "0910/0910.4842_arXiv.txt": { "abstract": "{ Laser Comb Wavelength calibration shows that the ThAr one is locally unreliable with possible deviations of up to 100 \\ms within one order range, while delivering an overall 1 \\ms accuracy (Wilken et al 2009). Such deviation corresponds to \\daa $\\approx 7\\cdot 10^{-6}$ for a Fe\\,{\\sc ii}-Mg\\,{\\sc ii} pair. Comparison of line shifts among the 5 Fe\\,{\\sc ii} lines, with almost identical sensitivity to fine structure constant changes, offers a clean way to directly test the presence of possible local wavelength calibration errors of whatever origin. We analyzed 5 absorption systems, with \\zabs ranging from 1.15 to 2.19 towards 3 bright QSOs. The results show that while some lines are aligned within 20 \\ms, others reveal large deviations reaching 200 \\ms or higher and corresponding to a \\daa $\\ge 10^{-5}$ level. The origin of these deviations is not clearly identified but could be related to the adaptation of wavelength calibration to CCD manufacturing irregularities. These results suggest that to draw conclusions from \\daa analysis based on one or only few lines must be done with extreme care. ", "introduction": "Investigations into possible systematic effects entering in the measurement of \\daa have been discussed among others in Murphy et al (2003), Chand et al (2006), Molaro et al (2008). Here we focus onto the wavelength calibration error. The wavelength calibration of the QSO CCD images is made by comparison with ThAr lamp taken before and after the QSO frames. The laboratory wavelengths used in the line identification have uncertainties between 10-150 \\ms (Palmer \\& Engleman 1983, Lovis \\& Pepe 2007 ). Errors in the wavelength calibration could be estimated from the mean wavelength-pixel residuals from the polynomial best solution. Residuals vary from 3-4 m\\AA (Chand et al 2006) to 1 m\\AA, or even better, (Levshakov et al 2006, Thompson et al 2009). Murphy et al (2003) treating the ThAr as if they were QSO lines obtained that the error in the individual absorption systems is typically few times $10^{-6}$ but there are significant cases up to $5 \\cdot 10^{-6}$ or higher. When averaged over a large sample of simulated systems they obtained \\daa = $ 0.4 \\pm 0.8\\cdot 10^{-7}$ and concluded that these errors are not expected to drive line shifts in the QSO absorption-line results. A similar conclusion was achieved by Chand et al (2006) where 3 and 2 ThAr lines were taken in proximity of FeII and MgII lines, respectively. However, the residuals in the wavelength calibration may not reflect the real wavelength error entirely. Recently an experiment conducted at HARPS with the first Laser Comb wavelength calibration in the optical revealed significant deviations of the ThAr wavelength calibration with peak-to-peak deviations of up to 100 \\ms within one echelle order. Similarly in HIRES-Keck observations a comparison of the ThAr calibration with a I2 self-calibration spectrum revealed significant and not reproducible deviations (Griest et al 2009) ", "conclusions": "By using line shifts measurements of the five Fe\\,{\\sc ii} lines with identical sensitivity to $\\alpha$ changes we have studied possible systematics in the wavelength calibration. We found that: \\begin{itemize} \\item{In most of the systems one or more lines are deviating by several hundreds of \\ms. This exceeds the likely errors estimated from the residuals and reveals the presence of hidden systematics in the spectral data. } \\item{ These errors could severely affect the \\daa measure, in particular in the methods where \\daa is derived from few lines or where some lines are crucial for the result. To have a reliable measure it is important to have consistent \\daa obtained from several lines for a given absorption system. } \\item{HE 0001-2340, which is one of the QSO used in the present UVES controversy, reveals very poor quality and the whole sample should be reanalyzed {\\it ab initio} } \\end{itemize}" }, "0910/0910.4904_arXiv.txt": { "abstract": "At energies $\\gtrsim$ 2 keV, active galactic nuclei (AGN) are the source of the cosmic X-ray background (CXB). For AGN population synthesis models to replicate the peak region of the CXB ($\\sim$30 keV), a highly obscured and therefore nearly invisible class of AGN, known as Compton thick (CT) AGN, must be assumed to contribute nearly a third of the CXB. In order to constrain the CT fraction of AGN and the CT number density we consider several hard X-ray AGN luminosity functions and the contribution of blazars to the CXB. Following the unified scheme, the radio AGN luminosity function is relativistically beamed to create a radio blazar luminosity function. An average blazar spectral energy density model is created to transform radio luminosity to X-ray luminosity. We find the blazar contribution to the CXB to be 12$\\%$ in the 0.5-2 keV band, 7.4$\\%$ in the 2-10 keV band, 8.9$\\%$ in the 15-55 keV band, and 100$\\%$ in the MeV region. When blazars are included in CXB synthesis models, CT AGN are predicted to be roughly one-third of obscured AGN, in contrast to the prediction of one half if blazars are not considered. Our model implies a BL Lac X-ray duty cycle of $\\sim$13$\\%$, consistent with the concept of intermittent jet activity in low power radio galaxies. ", "introduction": "\\label{sect:intro} Nearly half a century after the cosmic X-ray background (CXB) was discovered \\citep{g62}, the majority of the CXB up to 10 keV has been resolved into distinct point sources by deep observations conducted by {\\em ROSAT}, {\\em Chandra}, and {\\em XMM-Newton} \\citep{h98, m00, g01, g02, h01, a03, w04, bh05, w05}. These discrete sources are active galactic nuclei (AGN), compact extra-galactic sources powered by accretion onto black holes \\citep{lb69,r84}. As such, the CXB encapsulates the history of accretion onto super massive black holes and provides a powerful tool to aid scientific understanding of accretion processes \\citep{fb92}. It has been shown that a large portion of this accretion is shrouded from our view by intervening matter along the line of sight \\citep{sw89, cel92, m94, c95, f99, t09}. For AGN spectral and spatial density models to match the peak of the CXB at $\\sim$30 keV, the models must predict a large number of highly obscured sources known as Compton thick (CT) AGN \\citep{r99, u03, tu05, b06, gch07}, which have a neutral hydrogen column density $N_{\\mathrm{H}}\\gtrsim$ 1.5 $\\times$ 10$^{24}$ cm$^{-2}$ \\citep{t09}, making them practically invisible in the 2-10 keV band \\citep{g94}. CXB synthesis models predict CT sources make up roughly half of the obscured AGN population \\citep{r99, u03, tu05, b06, gch07}. In recent years, studies to observationally constrain the CT fraction have been undertaken. At first, small local studies seemed to agree with the model predictions that half of all obscured AGN are CT \\citep{r99, gua05}. However, a recent study by \\citet{t09}, using samples from {\\em INTEGRAL} and {\\em Swift} observations and high redshift, IR-selected CT AGN candidates, suggests that CT AGN contribute only about 9$\\%$ of the X-ray background, which contrasts sharply to the prediction by \\citet{gch07} that CT AGN account for nearly a third of the CXB. \\citet{m09} studied 88 AGN observed by INTEGRAL/IBIS in the 20-40 keV band and found that at least 16$\\%$ of obscured AGN are CT and $\\gtrsim$24$\\%$ of the AGN in their local sample ($z \\leq 0.015$) are CT. If $\\sim$75$\\%$ of local AGN are obscured \\citep{r99, t09}, the local AGN sample of \\citet{m09} suggests that $\\gtrsim$32$\\%$ of obscured AGN are CT. Given the uncertainity of the CT AGN fraction, smaller classes of CXB contributors must be considered. The CXB contribution from the small class of AGN known as blazars has previously been ignored by CXB synthesis models and CT AGN fraction predictions, even though blazars are known to emit in a broad range from radio to TeV energies. To further constrain model predictions of the CT AGN fraction, the blazar class of AGN must be considered. Blazars are a unique and extreme class of AGN. Unified models of AGN, as summarized by Antonucci (1993) and Urry \\& Padovani (1995), explain blazars as radio galaxies with relativistic jets viewed close to the line of sight. Flat spectrum radio quasars (FSRQs) are relativistically beamed FRIIs (luminous radio galaxies) and BL Lac objects (BL Lacs) are relativistically beamed FRIs (less luminous radio galaxies). The features which define blazars (extreme variability, high luminosity, high polarization, and radio core-dominance) are due to the relativistic beaming caused by looking down the relativistic jet of the blazar \\citep{p07b}. The details of the spectral energy distribution (SED) of blazars is still a topic riddled with uncertainties \\citep{km08}. The extreme variability that distinguishes the blazar class necessitates simultaneous multi-wavelength observations to understand the spectral properties \\citep{gt08}. However, the two-hump form of the blazar SED, spanning from radio to $\\gamma$-ray energies, is well known \\citep{u99}. The lower energy hump is due to synchrotron radiation while the higher energy hump is due to inverse Compton scattering \\citep{up95}. It has been shown that blazars are significant progenitors of the $\\gamma$-ray background \\citep{g06, nt06, km08}; therefore it is expected that blazars should have a non-negligible contribution to the CXB. \\citet{g06} predict that blazars should account for 11-12$\\%$ of the soft CXB around 1 keV; however, no estimation is made for the blazar contribution to the peak region of the CXB around 30 keV. A recent study by \\citet{a09}, based on the three year {\\em Swift}/BAT blazar sample, claims that blazars contribute about 10$\\%$ of the X-ray background in the 2-10 keV band. In the 15-55 keV band \\citet{a09} predict blazars contribute $\\sim$20$\\%$ if blazars are modeled as a single population or $\\sim$9$\\%$ if FSRQs and BL Lacs are modeled as two distinct populations. Both \\citet{g06} and \\citet{a09} found that blazars could contribute 100$\\%$ of the CXB in the MeV band. Due to uncertainties in the low luminosity end of the AGN hard X-ray luminosity function (HXLF), multiple HXLFs must be considered (e.g., Ueda et al. 2003; La Franca et al. 2005; Silverman et al. 2008; Aird et al. 2009; Ebrero et al. 2009; Yencho et al. 2009) to understand the range of predicted CT AGN. Recent AGN HXLFs find that luminosity-dependent density evolution (LDDE) provides the best fit to the observational data \\citep{u03, h05, lf05, si08, e09, y09}. \\citet{aird09} find a new evolutionary model, luminosity and density evolution (LADE), also fits the observational data well. Both the LDDE and LADE models are in keeping with the findings that the scarce, high-luminosity sources, quasars, show sharp positive evolution from $z$ $\\approx$0-2, while less luminous sources, Seyferts, evolve more temperately \\citep{b05, bh05, h05}. Given the connection between AGN and galaxy evolution (e.g., Ferrarese \\& Merritt 2000; Smol\\v{c}i\\'{c} 2009), it is not surprising that AGN evolution matches the trend of galaxy formation, where massive galaxies formed earlier in cosmological time while smaller structures have waited until more recent times to form (e.g., Cowie et al. 1999). In this work the blazar contribution to the CXB is predicted and the implications for the CT AGN fraction are discussed in the context of multiple HXLFs. In \\S \\ref{sect:calc} we present the model used for the blazar and non-blazar AGN contributions to the CXB. In \\S \\ref{sect:res} our results are presented while discussions and conclusions are given in \\S \\ref{sect:sum}. We assume a $\\Lambda$CDM cosmology with $H_0$ = 70 km s$^{-1}$ Mpc$^{-1}$, $\\Omega_{\\Lambda}$ = 0.7, and $\\Omega_{m}$ = 0.3 \\citep{s07}. ", "conclusions": "\\label{sect:sum} It is clear that blazars make a non-negligible contribution to the CXB and significantly reduce the number of CT AGN predicted, and may be primarily responsible for the MeV background. This paper presents an upper limit to blazar contribution to the CXB by utilizing the unified model of radio-loud AGN. The main conclusions found here do not change with a different choice of AGN radio luminosity function (e.g., Condon et al. 2002; Sadler et al. 2002; Best et al. 2005; Kaiser \\& Best 2007; Mauch \\& Sadler 2007); however, beaming parameters need to be modified as these luminosity functions do not distinguish between FRIs and FRIIs or high and low luminosity sources. A recent study by \\citet{a09}, using the three year sample of {\\em Swift}/BAT blazars, finds similar results for FSRQs as those found here; however, this work finds a greater contribution to the CXB and cosmic $\\gamma$-ray background by BL Lacs. Due to the small number statistics and small redshift range of the {\\em Swift}/BAT BL Lac sample, \\citet{a09} are not able to uniquely determine the evolutionary parameters and thus assume no evolution for BL Lacs. This work assumes BL Lacs evolve in the same manner as low luminosity radio galaxies. Also, \\citet{a09} assume a simple power law SED model for BL Lacs whereas this work utilizes an SED model based on average BL Lac properties taking into account the variety of BL Lac subclasses of LBLs and HBLs. Several studies have found that BL Lacs contribute substantially to the cosmic $\\gamma$-ray background \\citep{g06, nt06, km08}, thus it is not expected that the BL Lac contribution to the CXB is negligible, as found by \\citet{a09}. As this work uses a more physical BL Lac SED model and a reasonable evolutionary model, we expect that this work may more accurately model the BL Lac contribution to the CXB, although the factor 2.5 discrepancy in the BL Lac source counts in the hard X-ray band must be solved in future works. \\citet{t09} find the density of CT AGN at $z=0$ with $L_X$ $>$ 10$^{43}$ erg s$^{-1}$ is $\\sim$2.2 $\\times$ 10$^{-6}$ Mpc$^{-3}$. The luminosity function of \\citet{u03} predicts the density of CT AGN with $L_X$ $>$ 10$^{43}$ erg s$^{-1}$ at $z=0$ to be 7.3 $\\times$ 10$^{-6}$ Mpc$^{-3}$ if blazars are not considered and 4.4 $\\times$ 10$^{-6}$ Mpc$^{-3}$ if blazars are considered. With the blazar contribution to the CXB considered, the \\citet{u03} over predicts the CT AGN density found by \\citet{t09}, by a factor of 2. Conversely, the luminosity function proposed by \\citet{e09} predicts the density of CT AGN with $L_X$ $>$ 10$^{43}$ erg s$^{-1}$ at $z=0$ to be 1.1 $\\times$ 10$^{-6}$ Mpc$^{-3}$ if blazars are not considered and 3.6 $\\times$ 10$^{-7}$ Mpc$^{-3}$ if blazars are considered, which is a factor of 6 smaller than the density reported by \\citet{t09}. According to the INTEGRAL results of \\citet{m09}, the $f_{CT}$ $\\geq$ 0.32 with no upper limit given. Between different HXLFs there is a large scatter in the predicted $f_{CT}$ and the predicted CT AGN density varies by a factor of ~30. This clearly illustrates the limits imposed by the uncertainty of the low luminosity end of the AGN HXLF and how important it is for future missions to probe this portion of the HXLF. It has been shown that blazars, specifically BL Lacs, contribute the majority of the $\\gamma$-ray background \\citep{g06, nt06, km08}. \\citet{g06} found that unless BL Lacs have a small high energy duty cycle the predicted blazar $\\gamma$-ray emission would over predict the $\\gamma$-ray background. Furthermore, \\citet{km08} suggests that using radio blazar luminosity functions may cause an overestimation of the number of sources emitting robustly at higher energies, as it is not certain that all radio sources will have strong X-ray and $\\gamma$-ray emission. Physical and evolutionary models of quasars indicate that AGN activity is short-lived and possibly recurrent \\citep{s82, cp89, ct92}. \\citet{fran98} showed that long-lived, continuous AGN activity is not consistent with the black hole mass function they calculated from their sample of 13 local galaxies, but short-lived and recurrent AGN activity matches the data well. Several sources which appear to be restarted AGNs have been observed \\citep{b83, r93, s99, v04, j07, f09}. Sources have also been observed which have relic radio lobes but the AGN activity is not currently in an active phase \\citep{parm07, dk09, f09}. Recent observations by {\\em Hubble Space Telescope} and {\\em Chandra} of the relativistic jet of nearby M87 provide evidence for the intermittent nature of jet X-ray emission (e.g., Perlman et al. 2003; Harris et al. 2006; Stawarz et al. 2006; Madrid 2009). Large amplitude flaring has been observed from the previously quiescent knot HST-1 in the jet of M87 since 2000 \\citep{mad09}. This flaring activity is shown to be consistent with shocks occurring within the jet as faster moving particles collide with slower relativistic particles injected into the jet at an earlier time \\citep{perl03, sta06, mad09}. \\citet{perl03} and \\citet{sta06} suggest the recent X-ray flaring of HST-1 is directly related to material injected at the base of the jet 30-40 years ago. Therefore, a jet X-ray duty cycle is expected. Finally, due to the spectral steepening that occurs after the flow of energetic particles into the jet has ceased, the best frequency range to search for relic radio lobes is the low radio regime, less than 1 GHz \\citep{parm07}. Therefore, it is likely that the radio AGN luminosity function given by \\citet{w01} at 151 MHz includes relic radio lobes. As this would affect the low luminosity end of the luminosity function more prevalently as relic sources tend to not be as luminous as active sources \\citep{dk09}, the BL Lac luminosity function found here may overpredict the number of BL Lacs. Thus, the average BL Lac X-ray duty cycle is likely to be somewhat larger than the 13$\\%$ found here." }, "0910/0910.0021.txt": { "abstract": "We present results from multi-epoch spectral analysis of {\\sl XMM-Newton} and {\\sl Chandra} observations of the broad absorption line (BAL) quasar \\apm. Our analysis shows significant X-ray BALs in all epochs with rest-frame energies lying in the range of $\\sim$ 6.7--18~keV. The X-ray BALs and 0.2--10~keV continuum show significant variability on timescales as short as 3.3~days (proper time) %confirming the intrinsic nature of the absorbing outflow and implying a source size-scale of $\\sim$ 10~$r_{\\rm g}$, where $r_{\\rm g}$ is the gravitational radius. %The fitted width of the X-ray absorption troughs imply a We find a large gradient in the outflow velocity of the X-ray absorbers with projected outflow velocities of up to 0.76~$c$. %The detected projected maximum outflow velocity of $v \\sim 0.76~c$ The maximum outflow velocity constrains the angle between the wind velocity and our line of sight to be less than $\\sim$ 22$^{\\circ}$. Based on our spectral analysis we identify the following components of the outflow: (a) Highly ionized X-ray absorbing material with an ionization parameter in the range of $2.9 \\simlt \\log\\xi \\simlt 3.9$ (the units of $\\xi$ are erg~cm~s$^{-1}$) and a column density of $\\log{N_{\\rm H}} \\sim 23$ (the units of $N_{\\rm H}$ are cm$^{-2}$) outflowing at velocities of up to 0.76~$c$. (b) Low-ionization X-ray absorbing gas with %a column density of $\\log{N_{\\rm H}} \\sim 22.8$. We find a possible trend between the X-ray photon index and the maximum outflow velocity of the ionized %X-ray absorber in the sense that flatter %X-ray spectra appear to result in lower outflow velocities. %We provide a plausible explanation for this effect. Based on our spectral analysis of observations of \\apm\\ over a period of 1.2 years (proper time) we estimate the mass-outflow rate and efficiency of the outflow to have varied between $16_{-8}^{+12}$ $M_{\\odot}$~yr$^{-1}$ and $64_{-40}^{+66}$ $M_{\\odot}$~yr$^{-1}$ and $0.18_{-0.11}^{+0.15}$ to $1.7_{-1.2}^{+1.9}$, respectively. Assuming that the outflow properties of \\apm\\ are a common property of most quasars at similar redshifts, our results then imply that quasar winds are massive and energetic enough to influence significantly the formation of the host galaxy, provide significant metal enrichment to the interstellar medium (ISM) and intergalactic medium (IGM), and %ISM and IGM, and are a viable mechanism for feedback at redshifts near the peak in the number density of galaxy mergers. ", "introduction": "A number of physical mechanisms of interactions of AGNs with their environments have been proposed to explain feedback in galaxies and clusters of galaxies. These feedback mechanisms include radio jets, AGN winds, and AGN radiative heating. %, SN type 1 and starburst winds. It may be the case that several of these mechanisms operate simultaneously and their relative contribution to the feedback process varies depending in part on the size of the super-massive black hole (SMBH), and the combined gravitational potential of the SMBH and host galaxy. The gas and dust content of the host galaxy, the redshift and age of the host, and the inflow and outflow properties of the central super-massive black hole may be important as well. The radio-jet feedback mechanism was proposed after the discovery of significant radio cavities and shock fronts in several clusters of galaxies. The cavity sizes are large enough to imply that the amount of energy injected in the intracluster medium (ICM), as inferred from the mechanical work $pdV$ required to displace the gas in the radio cavities, is sufficient to balance radiative losses, suppress cooling flows, and halt star formation (e.g., Fabian et al. 2009, and references therein). %Assuming the cavities are inflated by mechanical energy from the radio jets it is found that %the ratio of mechanical energy to radio luminosity ranges by three orders of magnitude. The radio jets are thought to be driven by either accretion onto the black hole (e.g., Blandford \\& Payne 1982) or the black-hole spin (e.g., Blandford \\& Znajek 1977). The details of the process by which radio jets deposit energy and distribute it uniformly throughout the ICM is currently not well understood. Another possibly important mode that may contribute to feedback is AGN radiation. Radiative heating (e.g., Ciotti \\& Ostriker 2007) has been proposed as an alternative to the radio-jet mechanism to explain the suppression of cooling flows in isolated elliptical galaxies and possibly in clusters of galaxies. %SN type I and starbust winds are observed in many galaxies and expected to contribute %to the feedback process. %The contribution of SN type I and starburst winds to the feedback process %may be less important in massive galaxies and %clusters of galaxies where the gravitational potential %may be too large for the gas in these relatively low velocity winds to escape. %It is also important to note that most of the observations of SN type I and starbust winds %are of relatively nearby galaxies and little is known for redshifts above $z {\\simgt} 1.5$ where %the number density of mergers is expected to peak. In field galaxies, especially ones in the redshift range of $z \\approx 1-3 $ where the number density of galaxy mergers is thought to peak (e.g., Di Matteo et al. 2005), quasar winds are thought to be one of the major contributors to feedback. Winds can be driven by radiation pressure, magneto-centrifugal forces, and thermal pressure or a combination of these processes. The potential importance of quasar outflows has been explicitly demonstrated in theoretical models of structure formation and galaxy mergers that incorporate the effects of quasar outflows (e.g., Silk \\& Rees 1998; Scannapieco \\& Oh 2004; Granato et al. 2004; Springel, Di Matteo, \\& Hernquist 2005; Hopkins et al. 2005, 2006). Observational evidence to support these theories of quasar outflows was provided by the discovery of near-relativistic X-ray absorbing winds in several broad absorption line (BAL) quasars (e.g., Chartas et al. 2002, 2003) and later in several non-BAL quasars (e.g., Reeves et al. 2003; Pounds et al. 2003, 2006; Braito et al. 2007). The relative contribution of each of these proposed feedback mechanisms will vary in part depending on the following conditions: (a) The importance of the radio-jet mode will depend on the radio loudness and the duty cycle of the central AGN. The presence of a radio jet in an AGN is likely to depend on the available supply of infalling material to feed the black hole and/or the spin of the black hole. We note that several observations imply that the fraction of radio-loud objects varies with redshift and luminosity (e.g., Peacock et al. 1986; Schneider et al. 1992; La Franca et al. 1994; Jiang et al. 2007). Specifically it is found that the radio-loud fraction decreases with redshift and increases with luminosity. The radio jet mode therefore may be more important at later times and in clusters of galaxies and in nearby massive ellipticals that have a larger probability of containing a radio-loud object compared to galaxies at $ z {\\simgt} 1$ which are more likely to contain a radio-quiet AGN. %(b) A starburst may drive a galactic wind under the condition that the %kinetic energy injected into the disk exceeds its binding energy. %Starburst winds can therefore be an important feedback mechanism %in the case where the escape velocity is low, the %lifetime of the starburst is relatively long and the %mass density of the gas is high. (b) The feedback process is also expected to depend on the mass of the central black hole and the gravitational potential of the host galaxy. In order for a feedback mechanism to be effective it must be able to drive gas out of the host with a velocity that is larger than the escape velocity determined by the gravitational potential of the system. The mass outflow rates and outflow efficiencies of AGN winds will also depend strongly on the mass and luminosity of the SMBH. Analyses of the kinematic properties of UV absorbers in quasars indicate a strong increase in outflow velocity of UV absorbers with quasar UV luminosity (e.g., Laor \\& Brandt 2002; Ganguly \\& Brotherton 2008). Luminous quasars are found to contain winds with velocities of the UV outflowing material of up to 60,000~km~s$^{-1}$ whereas Seyfert galaxies typically have winds with significantly lower velocities of up to $\\sim$ 2,000 km~s$^{-1}$. %with UV outflowing absorbers of up to ~2000 km/s. (c) Hydrodynamic simulations of AGN atmospheres driven by radiation pressure indicate that the strengths of the outflows originating from AGN accretion disks depend on the Eddington ratio ($L_{\\rm Bol}/L_{\\rm Edd}$) with more massive and faster winds produced at larger Eddington ratios (e.g., Proga, Stone \\& Drew 1988; Proga, Stone \\& Kallman 2000; Proga \\& Kallman 2004). In systems with a limited supply of gas to fuel the AGN it is expected that AGN winds will be relatively weak. (d) In the case of a galaxy merger the degree to which each mechanism contributes to the overall feedback process may change with time since the merger event occurred. Observations in X-rays of the BAL quasar APM~08279+5255, the mini-BAL quasar PG~1115+080, and perhaps the low-ionization BAL quasar H~1413-117 have strongly suggested the presence of near-relativistic outflows of X-ray absorbing material in these objects (Chartas et al. 2002, 2003, 2007a, 2007b). Observations of these quasars in the optical and UV indicate outflow velocities and column densities that are more than an order of magnitude lower than those inferred from the X-ray spectra. These differences in outflow velocities and absorbing column densities suggest that X-ray BALs probe a highly ionized, massive and high-velocity component of the wind that is largely distinct from the absorbers detected in the optical and UV wavebands. Most of the mass and energy of quasar winds is likely carried by the outflowing X-ray absorbers, and observations in the X-ray band are therefore crucial for improving our understanding of the contribution of quasar winds to feedback in galaxies. This paper will focus on determining the significance of the wind in the $z = 3.91$ gravitationally lensed BAL quasar \\apm\\ in the feedback process during an epoch near the peak of the number density of mergers. Specifically, we present results from analyses of several \\xmm\\ and \\chandra\\ observations of \\apm\\ in order to determine the outflow velocity and its variability, obtain insights on the acceleration mechanism, constrain the geometry of the wind, and estimate the mass outflow rate and the outflow efficiency. Throughout this paper we adopt a $\\Lambda$-dominated cosmology with $H_{0}$ = 70~km~s$^{-1}$~Mpc$^{-1}$, $\\Omega_{\\rm \\Lambda}$ = 0.7, and $\\Omega_{\\rm M}$ = 0.3. ", "conclusions": "We have presented results from an analysis of three \\xmm\\ and two \\chandra\\ observations of the $z = 3.91$ BAL quasar \\apm. The main goals of these observations were to study the kinematic and photoionization properties of the wind in order to assess whether it plays an important role in controlling the evolution of the host galaxy and central black hole. Additional objectives included understanding the mechanism that drives the X-ray absorbers to near-relativistic velocities and constraining the geometry of the outflow. The main conclusions of our spectral and timing analyses are the following: (a) X-ray BALs lying in the range of 1.5--3.6~keV (observed frame) are detected at the greater than 99.9\\% confidence level in all five observations of \\apm. We note that a recent analysis of three {\\sl Suzaku} observations of \\apm\\ also shows significant detections of X-ray BALs in all three observations (Saez et al. 2009). (b) Based on our spectral analysis we identify the following components of the outflow. First, a highly ionized X-ray absorbing material with an ionization parameter in the range of $2.9 \\simlt \\log\\xi \\simlt 3.9$ and a column density of $\\log{N_{\\rm H}} \\sim 23$ outflowing at velocities of up to 0.76~$c$, and second a low-ionization X-ray absorbing gas with a column density of $\\log{N_{\\rm H}} \\sim 22.8$ that may constitutes the X-ray shielding gas that is thought to protect the UV absorber from becoming over-ionized. Models that include partial covering result in even larger column densities of the ionized absorber of up to $\\log{N_{\\rm H}} \\sim 24$, however, we caution that these values are not well constrained (see model 9 of Table~3) by the spectral fits. We note that no iron overabundance is required to fit the X-ray spectra of \\apm\\ consistent with the spectral analysis of previous observations of this object. (c) Significant variability of the X-ray continuum and X-ray BALs over short and long time-scales is detected. Specifically, we detect a $\\sim$ 36.8 $\\pm$ 0.3 \\% change of the 0.2--10~keV pn count-rate between epochs 3 and 4 that are separated by only 3.3~days (proper-time). Based on a simple light-travel time argument this variability time-scale implies a radius of the emission region of $\\sim$ 7.4 $\\times$ 10$^{15}$~cm which is comparable to $r_{\\rm ISCO}$ = 4.5 $\\times$ 10$^{15}$~cm for the case of a Schwarzschild black hole in \\apm. We speculate based on the energy independent spectral change between epochs 3 and 4 (with the exception of the Fe BAL region) that the cause of this variability is a change in the covering fraction as the outflowing absorbing material moves across our line of sight. Significant changes in the flux densities between epochs 2 and 3 that are separated by $\\sim$ 5.4 years (observed-frame) are detected. The variation of flux density is significant below 3~keV (observed frame) showing a increase toward lower energies but is near zero above 3~keV. This type of energy-dependent change suggests that the intrinsic unabsorbed luminosity of \\apm\\ has not significantly varied between these epochs and the long-term variability is the result of a decrease in the opacity of the absorber. This interpretation is consistent with the results of our spectral analysis. (d) Assuming our interpretation that the high-energy X-ray BALs are produced by resonance transitions of highly-ionized iron %Fe XXV(1s--2p) we estimate the mass-outflow rate and efficiency of the outflow to have varied over a period of 1.2 years (proper-time) between $16_{-8}^{+12}$ $M_{\\odot}$~yr$^{-1}$ and $64_{-40}^{+66}$ $M_{\\odot}$~yr$^{-1}$ and $0.18_{-0.11}^{+0.15}$ to $1.7_{-1.2}^{+1.9}$, respectively. (e) The detected projected maximum outflow velocity of $v_{\\rm max} \\sim 0.76c$ leads to a constraint on the angle between the outflow direction and the observed line of sight through the absorber of $\\theta_{\\rm max}$ $\\simlt$ 22$^{\\circ}$ (assuming \\FeXXV($1s^2-1s2p$)) and $\\theta_{\\rm max}$ $\\simlt$ 26$^{\\circ}$ (assuming \\FeXXVI($1s-2p$)). Such a small angle is consistent with the unification scheme of BAL and non-BAL quasars that posits that BAL quasars are viewed almost along the outflow direction. (f) A possible trend is found between the X-ray photon index and the maximum outflow velocity of the ionized X-ray absorber in the sense that flatter X-ray spectra appear to result in lower outflow velocities. One possible explanation is that flatter X-ray spectra over-ionize the X-ray absorber resulting in a decrease of the force multiplier of the X-ray absorber and thus a lower outflow velocity." }, "0910/0910.2305_arXiv.txt": { "abstract": "We present FLAMES/GIRAFFE spectroscopy obtained at the Very Large Telescope (VLT). Using these observations we have been able for the first time to observe the Li I doublet in the Main Sequence stars of a Globular Cluster. We also observed Li in a sample of Sub-Giant stars of the same B-V colour. Our final sample is composed of 84 SG stars and 79 MS stars. In spite of the fact that SG and MS span the same temperature range we find that the equivalent widths of the Li I doublet in SG stars are systematically larger than those in MS stars, suggesting a higher Li content among SG stars. This is confirmed by our quantitative analysis which makes use of both 1D and 3D model atmospheres. We find that SG stars show, on average, a Li abundance higher by 0.1\\,dex than MS stars. We also detect a positive slope of Li abundance with effective temperature, the higher the temperature the higher the Li abundance, both for SG and MS stars, although the slope is slightly steeper for MS stars. These results provide an unambiguous evidence that the Li abundance changes with evolutionary status. The physical mechanisms that contribute to this are not yet clear, since none of the proposed models seems to describe accurately the observations. Whether such mechanism can explain the cosmological lithium problem, is still an open question. ", "introduction": "The primordial Li abundance inferred from the fluctuations of cosmic microwave background measured by the WMAP satellite (\\cite[Spergel et al. 2007, Cyburt et al. 2008]{spe07,cyburt}) is $\\log ({\\rm Li}/{\\rm H})+ 12 =2.72\\pm0.06$, approximately 0.3--0.5 dex higher than the Li abundance determined in metal-poor stars of the Galactic halo. Many studies have tried to explain this difference: (a) \\cite[Piau et al. 2006]{pia06} propose that the first generation of stars, Population III stars, could have processed some fraction of the halo gas, lowering the lithium abundance; (b) other authors suggest that the primordial Li abundance has been uniformly depleted in the atmospheres of metal-poor dwarfs by some physical mechanism (e.g. turbulent diffusion as in \\cite{ric05,kor06}; gravitational waves as in \\cite{cat05}, etc.); and (c) finally, it has been also suggested that the standard Big Bang nucleosynthesis (SBBN) calculations should be revised, possibly with the introduction of new physics as in e.g. \\cite{jed04,jed06,jittoh,hisano}. Here we present the Li abundances of subgiant (SG) and main-sequence (MS) stars of the cluster NGC~6397. This study represents the first Li abundance measurements in MS stars of a Globular Cluster. ", "conclusions": "In Fig.~\\ref{figali} we display the derived 1D-NLTE Li abundances versus the 1D effective temperatures of dwarf and subgiant stars of the globular cluster NGC 6397 (see Fig.~2 in \\cite{gon09} for a similar picture but with $T_{\\rm eff}$ and Li abundances computed using 3D models). The stars have been divided into five bins of effective temperature. The error bar in the Li abundance is the dispersion for each bin, divided by the square root of the number of stars in the bin. The Li abundance decreases with decreasing temperature. This lithium abundance pattern is different from what is found among field stars (\\cite[Mel\\'endez \\& Ram{\\'\\i}rez 2004]{mel04}, \\cite[Bonifacio et al. 2007]{bon07}, \\cite[Gonz\\'alez Hern\\'andez et al. 2008]{gon08}). The difference in the Li abundance of dwarfs and subgiants is $\\sim0.14$~dex if one computes the mean 1D A(Li) and the standard deviation of the mean for the two samples. For subgiants we find mean 1D Li abundances of $2.31\\pm0.01$, while for dwarfs $2.17\\pm0.01$. The 3D models provide similar results although slightly higher Li abundances. \\cite{lin09} also find such difference between the mean Li abundances in MSs and SGs, but only by 0.03~dex although still significant at 1$\\sigma$. However, this result is partly affected by the very narrow range of $T_{\\rm eff}$ for MSs deduced by \\cite{lin09} ($\\sim 80$~K) compared to the wide range ($\\sim 450$~K) for the SGs (see Fig.~7 online in \\cite[Gonz\\'alez Hern\\'andez et al. 2009]{gon09}). Our results imply that the Li surface abundance depends on the evolutionary status of the star. In Fig.~\\ref{figali} we show the Li isochrones for different turbulent diffusion models (\\cite[Richard et al. 2005]{ric05}). These models have been shifted up by 0.14 dex in Li abundance to make the initial abundance of the models, $\\log ({\\rm Li}/{\\rm H})=2.58$, coincide with the primordial Li abundance predicted from fluctuations of the microwave background measured by the WMAP satellite (\\cite[Cyburt et al. 2008]{cyburt}). The models assuming pure atomic diffusion, and, among those including turbulent mixing, T6.0 and T6.09, are ruled out by our observations. All such models predict that in dwarf stars Li should be either more abundant or the same as in subgiant stars. The only model which predicts a Li pattern which is qualitatively similar to that observed, is the T6.25 model. Models including atomic diffusion and tachocline mixing (\\cite[Piau 2008]{pia08}) do not seem to reproduce our observations, since they provide a constant Li abundance up to 5500\\,K. The sophisticated models which besides diffusion and rotation also take into account the effect of internal gravity waves (\\cite[Talon \\& Charbonnel 2004]{tac04}), seem to predict accurately the Li abundance pattern in solar-type stars, at solar metallicity (\\cite[Charbonnel \\& Talon 2005]{cat05}), but models at low metallicity are still needed. The cosmological lithium discrepancy still needs to be solved. Given that none of the existing models of Li evolution in stellar atmospheres matches the observations, we hope our results will prompt further new theoretical investigations." }, "0910/0910.5713_arXiv.txt": { "abstract": "{ We present an analysis of the Sommerfeld enhancement to dark matter annihilation in the presence of an excited state, where the interaction inducing the enhancement is purely off-diagonal, such as in models of exciting or inelastic dark matter. We derive a simple and accurate semi-analytic approximation for the $s$-wave enhancement, which is valid provided the mass splitting between the ground and excited states is not too large, and discuss the cutoff of the enhancement for large mass splittings. We reproduce previously derived results in the appropriate limits, and demonstrate excellent agreement with numerical calculations of the enhancement. We show that the presence of an excited state leads to generically larger values of the Sommerfeld enhancement, larger resonances, and shifting of the resonances to lower mediator masses. Furthermore, in the presence of a mass splitting the enhancement is no longer a monotonic function of velocity: the enhancement where the kinetic energy is close to that required to excite the higher state can be up to twice as large as the enhancement at zero velocity.} ", "introduction": "In particle physics, perturbation theory is commonly employed to calculate annihilation and scattering cross sections, with higher-order terms in the perturbative expansion being neglected. Provided the theory is not strongly coupled, this is generally a good approximation for relativistic particles, but at low velocities and in the presence of a long-range force (classically, when the potential energy due to the long-range force is comparable to the particles' kinetic energy), the perturbative approach breaks down. In the nonrelativistic limit, the question of how the long-range potential modifies the cross section for short-range interactions can be formulated as a scattering problem in quantum mechanics, with significant modifications to the cross sections occuring when the particle wavefunctions are no longer well approximated by plane waves (so the Born expansion is not well-behaved). The deformation of the wavefunctions due to a Coulomb potential was calculated by Sommerfeld in \\cite{sommerfeld}, yielding a $\\sim 1/v$ enhancement to the cross section for short-range interactions (where the long-range behavior due to the potential can be factorized from the relevant short-range behavior). Recent measurements of the positron and electron cosmic ray spectra by the PAMELA, ATIC, Fermi and H.E.S.S experiments have observed a rise in the positron fraction starting at $E \\sim 10$ GeV and extending (at least) up to $E \\sim 100$ GeV \\cite{Adriani:2008zr}, as well as a broad excess in the total $e^+ + e^-$ spectrum extending from several hundred GeV to $\\mathcal{O}(\\mathrm{TeV})$ energies \\cite{aticlatest, Abdo:2009zk, Aharonian:2009ah}. Observations by WMAP \\cite{Bennett:2003bz,Hinshaw:2008kr} also suggest an excess in microwave emission from the inner Galaxy, termed the ``WMAP Haze'' \\cite{Finkbeiner:2003im}, which has been attributed to synchrotron radiation from a new population of high energy electrons (with energies of tens to hundreds of GeV) \\cite{Finkbeiner:2004us, Dobler:2007wv, Hooper:2007kb}. There has been considerable interest in these observations as possible signatures of dark matter (DM) annihilation or decay. If DM annihilation is responsible for the excesses, the annihilation cross section must exceed that required for a WIMP initially in thermal equilibrium to freeze out with the correct relic density, by 1-4 orders of magnitude depending on the dark matter mass. Furthermore, to avoid antiproton bounds from the PAMELA experiment \\cite{Adriani:2008zq} and to generate sufficiently hard $e^+ e^-$ spectra, the dark matter should annihilate largely into leptons, pions, kaons and other light states (e.g. \\cite{Cirelli:2008pk, Meade:2009iu}). If dark matter is self-interacting, either via exchange of Standard Model gauge bosons or due to some novel interaction, then the nonperturbative ``Sommerfeld enhancement'' \\emph{must} be taken into account when computing annihilation cross sections in the present-day Galactic halo. The importance of the Sommerfeld enhancement in the context of dark matter annihilation has been studied by several authors \\cite{Hisano:2003ec, Hisano:2004ds, Cirelli:2007xd, MarchRussell:2008yu}. Due to the velocity dependence of the Sommerfeld effect, and the presence of large resonances due to near-threshold bound states at low velocity (see e.g. \\cite{MarchRussell:2008tu}), the annihilation cross section could potentially be greatly increased in the present-day Galactic halo ($\\beta \\sim 10^{-4}-10^{-3}$), without greatly perturbing the freezeout of annihilations at $\\beta \\sim 0.3$ (however, the Sommerfeld enhancement can modify the relic density at some level, particularly in the presence of resonances, see e.g. \\cite{Dent:2009bv, Zavala:2009mi}). Furthermore, it was pointed out in \\cite{Finkbeiner:2007kk,Cholis:2008vb} that the presence of light ($\\lesssim$ GeV) force carriers coupling to the dark matter would naturally lead to unusual dark matter annihilation signatures, with zero or small branching ratios into heavy hadrons and gauge bosons, and hard spectra of leptons, pions and other light states. The dominant annihilation channel in such models would be annihilation of the dark matter into the new force carriers, which would then decay into kinematically accessible SM states. In particular, provided the mass of the new force carrier was less than twice the proton mass, no excess antiprotons would be produced. A new GeV-scale force in the dark sector could therefore naturally generate both a large annihilation cross section via Sommerfeld enhancement, and the observed $e^+ e^-$ spectra, without violating antiproton constraints \\cite{ArkaniHamed:2008qn, Pospelov:2008jd}. As noted in \\cite{ArkaniHamed:2008qn}, if dark matter is charged under some new dark gauge group, then when the new gauge boson acquires an $\\mathcal{O}$(GeV) mass the states in the dark matter multiplet can naturally be split from each other by a small ($\\mathcal{O}$(MeV)) amount (this happens automatically, due to loops of the new dark gauge bosons, if the gauge group is non-Abelian \\cite{ArkaniHamed:2008qn, Baumgart:2009tn}, but such mass splittings can also be naturally generated in Abelian models, e.g. \\cite{Cheung:2009qd}). Furthermore, if the dark matter is a Majorana fermion or real scalar, its couplings with any vector boson must be off-diagonal, since it cannot carry conserved charge. Models of this type, where the dark matter can only scatter inelastically at tree level (that is, by exciting some slightly more massive state) are also motivated by the iDM \\cite{Smith:2001hy, Chang:2008gd} and XDM \\cite{Finkbeiner:2007kk} scenarios, which explain, respectively, the anomalies observed by the DAMA/LIBRA \\cite{Bernabei:2008yi} and INTEGRAL/SPI \\cite{Weidenspointner:2006nu} experiments. Despite the fairly generic presence of these excited states, most recent analyses of the Sommerfeld enhancement in the context of the cosmic-ray anomalies have treated only the case of degenerate dark matter states, under the presumption that the introduction of an excited state does not significantly modify the enhancement to annihilation, unless the mass splitting is large enough to cause the enhancement to cut off (see \\cite{ArkaniHamed:2008qn} for an argument to this effect) \\footnote{The related problem of collisional excitation of these states, via scattering mediated by a long-range force, has been considered by several authors, e.g. \\cite{Finkbeiner:2007kk,Chen:2009dm}.}. While it is true that the general qualitative features of the enhancement are mostly unchanged unless the mass splitting is quite large, modifications to the resonance structure occur for smaller mass splittings, and should be taken into account in any detailed analysis of the enhancement. However, the addition of even a single mass splitting adds an extra parameter to the problem, thus making numerical studies significantly more time-consuming; the numerical calculation of the enhancement can also become badly unstable in some parts of parameter space. In \\S \\ref{sec:deriv} we derive simple expressions to aid in computing the Sommerfeld enhancement in the general case where the $l$th partial wave dominates annihilation, for an arbitrary potential coupling $N$ states. In \\S \\ref{sec:parametrics} we discuss the general properties of the case where the dark matter has a single excited state and the interaction is purely off-diagonal, including the limit of zero mass splitting and a characterization of the parts of parameter space where the enhancement is negligible. In \\S \\ref{sec:approxsol} we derive an approximate semi-analytic solution for the Schr\\\"{o}dinger equation in the case of a two-state system with purely off-diagonal interaction, focusing on the case of $s$-wave annihilation. In \\S \\ref{sec:analysis} we present a simple analytic approximation for the Sommerfeld enhancement, strictly valid when the annihilation channels and matrix elements are the same for two particles in the excited state as in the ground state, and also applicable to the more general case (up to a simple rescaling), provided the enhancement is large. We employ this approximation to discuss the new features present in the case with a non-zero mass splitting, and confirm our results by numerical simulations. ", "conclusions": "We have developed a detailed approximate expression for the $s$-wave Sommerfeld enhancement in the presence of a mass splitting, in the case where only inelastic scattering between the interacting particles is possible at tree level. The $\\delta \\rightarrow 0$ limit of our results yields an accurate approximation for the enhancement due to an attractive scalar Yukawa potential. The enhancement cuts off at low velocity if the mass splitting $\\delta \\gtrsim \\alpha^2 m_\\chi$, the effective Bohr energy of the two-body state, but even for smaller mass splittings, interesting modifications of the enhancement relative to the $\\delta \\rightarrow 0$ case can occur. Additionally, we have derived simple expressions to facilitate computation of the Sommerfeld enhancement due to an arbitrary $N \\times N$ matrix potential, where the $l$th partial wave dominates. In the special case where all elements of the annihilation matrix are identical, our expression for the $s$-wave enhancement takes a simple form, \\[ S = \\frac{2 \\pi }{\\epsilon_v}\\sinh\\left(\\frac{\\epsilon_v \\pi }{\\mu}\\right) \\left\\{ \\begin{array}{cc} \\frac{1}{\\cosh\\left(\\epsilon_v \\pi/\\mu\\right)-\\cos\\left(\\sqrt{\\epsilon_\\delta^2-\\epsilon_v^2} \\pi /\\mu+2 \\theta_-\\right)} & \\quad \\epsilon_v < \\epsilon_\\delta, \\\\ \\\\ \\frac{\\cosh\\left(\\left(\\epsilon_v+\\sqrt{-\\epsilon_\\delta^2+\\epsilon_v^2}\\right) \\pi/2 \\mu\\right) \\text{sech}\\left(\\left(\\epsilon_v-\\sqrt{-\\epsilon_\\delta^2+\\epsilon_v^2}\\right) \\pi/2 \\mu \\right)}{ \\cosh\\left(\\left(\\epsilon_v+\\sqrt{-\\epsilon_\\delta^2+\\epsilon_v^2}\\right) \\pi /\\mu\\right)-\\cos(2 \\theta_-)} &\\quad \\epsilon_v > \\epsilon_\\delta. \\end{array} \\right. \\] where $\\mu$ and $\\theta_-$ are real functions which we have defined. This expression is valid provided $m_\\phi \\lesssim \\alpha m_\\chi$, $v/c \\lesssim \\alpha$, and $\\delta \\lesssim \\alpha^2 m_\\chi$, although it may become less accurate for $\\delta \\gtrsim \\max(\\alpha m_\\phi, \\, m_\\chi (v/c)^2)$. If any of the three former conditions are violated, there is no significant enhancement ($S \\sim \\mathcal{O}(1)$). In the regions where this expression is expected to be valid, comparison to numerical results demonstrates that it accurately reproduces all significant features of the enhancement. Furthermore, while derived for the case where all elements of the annihilation matrix are identical, a simple rescaling of this result by the average of the elements of the annihilation matrix (normalized to the $\\Gamma_{11}$ element) can accurately describe the enhancement for a general annihilation matrix. Employing this approximate form for the enhancement, we have shown that below the threshold for on-shell production of the excited state, in comparison to the case with no mass splitting: \\begin{itemize} \\item The non-resonant, unsaturated enhancement is higher by a factor of $\\sim 2$. \\item The positions of the resonances shift to lower values of $\\epsilon_\\phi$ ($m_\\phi / \\alpha m_\\chi$) and become more widely spaced. The heights of the resonances increase by a factor of $\\sim 4$. \\item The enhancement $S$ is no longer a monotonic function of velocity: close to the threshold for on-shell excitation of the higher-mass state, the enhancement may be as much as a factor of $\\sim 2$ greater than its saturated value (i.e. its value in the limit $v \\rightarrow 0$). \\end{itemize} In the context of DM annihilation, if the kinetic energy of WIMPs in the neighborhood of the Earth is close to the threshold for excitation of the higher-mass state (which is true by construction for models which realize the iDM mechanism), increased near-threshold enhancement may slightly alleviate constraints on DM annihilation from the early universe, where the DM is extremely slow-moving. The generically larger values of the unsaturated non-resonant enhancement, and the shift in position of the resonances, may help provide sufficiently large enhancements to generate recently observed cosmic-ray excesses, particularly for lower-mass force carriers. \\vskip 18pt \\noindent {\\bf Acknowledgements}: I am grateful to Nima Arkani-Hamed for suggesting this analysis, to Douglas Finkbeiner for many helpful discussions, and to Neal Weiner for pointing out useful references and the possibility of near-threshold resonances. I thank the anonymous referee for their useful feedback and suggestions; Lisa Goodenough, Rob Morris, David Poland, David Rosengarten, David Shih, Natalia Toro and Kathryn Zurek for helpful comments and conversations; and the NYU Center for Cosmology and Particle Physics and the Michigan Center for Theoretical Physics for their hospitality during the completion of this work. This work was partially supported by a Sir Keith Murdoch Fellowship from the American Australian Association. \\appendix" }, "0910/0910.0246_arXiv.txt": { "abstract": "This paper presents the results of collisional evolution calculations for the Kuiper belt starting from an initial size distribution similar to that produced by accretion simulations of that region - a steep power-law large object size distribution that breaks to a shallower slope at $r\\sim1-2$ km, with collisional equilibrium achieved for objects $r\\lesssim 0.5$ km. We find that the break from the steep large object power-law causes a divot, or depletion of objects at $r\\sim10-20$ km, which in-turn greatly reduces the disruption rate of objects with $r\\gtrsim 25-50$ km, preserving the steep power-law behavior for objects at this size. Our calculations demonstrate that the roll-over observed in the Kuiper belt size distribution is naturally explained as an edge of a divot in the size distribution; the radius at which the size distribution transitions away from the power-law, and the shape of the divot from our simulations are consistent with the size of the observed roll-over, and size distribution for smaller bodies. Both the kink radius and the radius of the divot center depend on the strength scaling law in the gravity regime for Kuiper belt objects. These simulations suggest that the sky density of $r\\sim1$ km objects is $\\sim10^6-10^7$ objects per square degree. A detection of the divot in the size distribution would provide a measure of the strength of large Kuiper belt objects, and constrain the shape of the size distribution at the end of accretion in the Kuiper belt. ", "introduction": "The Kuiper belt is a population of planetesimals outside the orbit of Neptune which exhibits high inclinations and eccentricities by many belt members \\citep{Trujillo2001b,Brown2001}, and has a very low mass, $M \\lesssim 0.1\\mbox{ M$_\\oplus$}$ \\citep{Gladman2001,Bernstein2004,Fuentes2008}. The high relative encounter velocities, $v_{rel} \\sim 1 \\mbox{ km s$^{-1}$}$ \\citep{Delloro2001} and infrequent collisions of the largest members \\citep{Davis1997,Durda2000} make the growth of Eris-sized bodies impossible over the age of the Solar system. Accretion in the early stages of planet-building must have been in a more dense and quiescent environment allowing large objects to grow \\citep{Stern1997a}. The size distribution of the Kuiper belt is the result of the accretionary and collisional processes that have gone on in that region, and therefore provides one of the main constraints on those processes. For a general review of the Kuiper belt size distribution, see \\citet{Petit2008} \\footnote{Much of the observations that constrain the size distribution for radii $r\\sim20-60$ km are published in more recent works \\citep{Fraser2009,Fuentes2009}.} There are three properties which describe the general shape of the Kuiper belt size distribution: \\begin{itemize} \\item the existence of the largest objects, Eris and Pluto \\item the large object size distribution for $D\\gtrsim100$ km which is well characterized by a steep power-law $\\frac{dN}{dR}\\propto r^{-q}$ with $q\\sim4.8$ \\citep{Gladman2001,Petit2006,Fraser2008} \\item the existence of a ``roll-over'' at $r\\sim25-60$ km where the size distribution flattens to a much shallower distribution than for larger objects \\citep{Bernstein2004,Fuentes2008,Fraser2009} \\end{itemize} \\noindent These three properties need to be reproduced by any successful attempt at recreating the accretionary and collisional history of the Kuiper belt region. While this has not yet been achieved, these properties provide some insight into the general history of the belt. Accretion simulations such as those of \\citet{Stern1996a,Kenyon2001} and \\citet{Kenyon2002} have demonstrated that objects as large as Eris could have accreted on timescales shorter than the age of the Solar system if the mass of the Kuiper belt was at least two orders of magnitude larger than the current mass, and if the relative encounter velocities, and hence, the eccentricities and inclinations of the objects were significantly lower in the early Solar system (see review by \\citet{Kenyon2008}). The existence of the steep large object size distribution however, suggests that accretion was a short-lived process \\citep{Kenyon2002,Fraser2008}. Some event must have disrupted accretion - likely the same event which scattered the primordial Kuiper belt onto orbits with high inclinations and eccentricities as observed today (see a review of proposed processes by \\citet{Morbidelli2008}) - otherwise the large object size distribution would be too shallow to be compatible with the observed distribution. The simulations of \\citet{Stern1996a,Kenyon2001} and \\citet{Kenyon2002} cannot reproduce the large roll-over size. In these simulations, interaction velocities remain low, such that objects larger than $r\\sim 1$ km do not experience disruptive collisions over the age of the Solar system. Thus, the size distribution for objects larger than $r\\sim 2$ km exhibits a steep slope comparable to that observed today, and the accretion break, or roll-over, occurs at radii too small to be compatible with observations. This suggests that after the belt was dynamically excited, it remained massive enough for a time long enough to allow collisional erosion to produce the large roll-over observed today. A model presented by \\citet{Kenyon2004} could reproduce the large roll-over if gravitational stirring by Neptune at its current location was included early in the simulations. Accretion occurred at a sufficient rate in these simulations to produce Eris-sized objects. In these simulations, accretion lasted too long however, resulting in a large object size distribution too shallow to be compatible with observations. In addition, the mass loss due to collisional grinding was insufficient to produce the tenuous modern-day belt (see discussions by \\citet{Morbidelli2008} and \\citet{Kenyon2008}). The Kuiper belt must have undergone a more rapid excitation and mass depletion than occurred in these calculations. \\citet{Pan2005} presented an analytical, order of magnitude, collisional evolution model, reminiscent of the ground breaking work of \\citet{Dohnanyi1969}. With this model, they demonstrated that the large roll-over size could be produced by collisional grinding on timescales shorter than the age of the Solar system. They however, assumed that the size distribution rolled over to collisional equilibrium. It has been demonstrated that there is a range of objects with sizes smaller than the roll-over, which are preferentially eroded but which are not being replenished from fragments of larger disrupted objects \\citep{Kenyon2002}. These objects do not achieve collisional equilibrium. Equilibrium is only achieved at a some smaller radius, the exact size of which, depends on the density of planetesimals. Thus, the model of \\citet{Pan2005} likely, does not predict the correct shape of the size distribution smaller than the roll-over, or the rate at which the roll-over evolves to larger sizes. \\citet{Benavidez2009} present a collisional evolution model of the Kuiper belt, in which they include collisions from multiple dynamical classes. Using this model, they calculate the collisional evolution of the Kuiper belt, starting from an initially steep size distribution for all sizes, with slope similar to that observed for large objects. Their results confirm the findings of \\citet{Pan2005}; collisional grinding with relative velocities comparable to that observed can disrupt the majority of objects with $r\\sim50-100$ km, and in effect, produce a roll-over at that size. These calculations however, likely do not reproduce the collisional history of the Kuiper belt, as they start from an initial condition not expected from standard accretion processes \\citep{Kenyon2002}. Accretion simulations which produce Eris-sized objects produce size distributions with an accretion break at $r\\sim 1-2$ km, not a steep distribution for all sizes. Here we present a collisional evolution model, which we utilize to calculate the size distribution of the Kuiper belt from some starting condition. With this model, we wish to determine whether or not collisional erosion could produce the $r\\sim20-50$ km roll-over, starting from a size distribution similar to that produced by models of accretion in the outer Solar system, but in the dynamically excited conditions of the current day Kuiper belt. From these calculations, we wish to constrain the shape of the modern day size distribution smaller than the roll-over, and to constrain the mass of the Kuiper belt at the end of accretion. In Section~\\ref{sec:model} we present our collisional evolution model, and the planetesimal strength and shattering models we adopt. In Section~\\ref{sec:results} we present the main results of our calculation, namely that that existence of a break at $r\\sim1$ km left-over from accretion causes a divot in the size distribution at larger sizes. In Section~\\ref{sec:discussion} we present the consequences of our model, and finish with concluding remarks in Section~\\ref{sec:conclusions}. ", "conclusions": "} Our calculations have provided the first direct link from the ``typical'' size distribution produced from accretion, to the collisionally modified size distribution of the modern day Kuiper belt. Our calculations have demonstrated that, given a size distribution similar to that produced by accretion, the likely result of collisional evolution in the Kuiper belt, is to produce a divot in the Kuiper belt size distribution at $r\\sim10-15$ km, and a roll-over from the large object accretion slope at $r\\sim25-60$ km, compatible with the observed roll-over in the Kuiper belt size distribution. We conclude that the size distribution is likely not a power-law for objects smaller than the roll-over, and exhibits a dearth of objects at the divot location. Assuming a divot has been formed, then a measurement of the divot center radius would provide constraint on the strength scaling law in the gravity regime for KBOs, and constrain the size distribution of $r\\sim 1$ km objects left-over from the accretion processes in the early Solar system. From our calculations, we find that the Kuiper belt must have had at least $1-5 \\mbox{ M$_\\oplus$}$ of material after its objects were excited onto orbits with eccentricities comparable to those observed, otherwise the roll-over would not be produced on Gyr timescales. If the belt had a mass of $\\sim 45 \\mbox{ M$_\\oplus$}$, then a divot would form in $\\gtrsim 40$ Myr. From the results of our calculations, we predict a deep absence of objects at $r\\sim10$ km. We also predict that the sky density of objects with $r\\sim1$ km is $\\sim10^6-10^7$ per square degree, 2-3 orders of magnitude larger than if the size distribution is a shallow power-law for objects smaller than the roll-over, as suggested by \\citet{Fraser2009}." }, "0910/0910.4086.txt": { "abstract": "A new classification system for carbon-rich stars is presented based on an analysis of 51 AGB carbon stars through the most relevant classifying indices available \\citep{Keenan93,Morgan04}. The extension incorporated, that also represents the major advantage of this new system, is the combination of the usual optical indices that describe the photospheres of the objects, with new infrared ones, which allow an interpretation of the circumstellar environment of the carbon-rich stars. This new system is presented with the usual spectral subclasses and \\CC -, j-, MS- and temperature indices, and also with the new SiC- (SiC/C.A. abundance estimation) and $\\tau$- (opacity) indices. The values for the infrared indices were carried out through a Monte Carlo simulation of the radiative transfer in the circumstellar envelopes of the stars. The full set of indices, when applied to our sample, resulted in a more efficient system of classification, since an examination in a wide spectral range allows us to obtain a complete scenario for carbon stars. ", "introduction": "The asymptotic giant branch stars with a ratio C$/$O $> 1$ have their optical spectra ruled by bands of carbon compounds, which obscure many atomic features. The green-red optical spectrum is dominated by Swan bands $^{12}$C$^{12}$C, Red System bands $^{12}$C$^{14}$N and sometimes present isotopic bands, e.g. $^{13}$C$^{12}$C and $^{13}$C$^{14}$N. As a result, a classification carried out through the optical atomic data on carbon stars is troublesome. During their ascent on the AGB phase, the mass loss from the star creates a circumstellar envelope of gas and dust. The compounds of this shell have their maximum emission on the infrared spectral region. The main infrared feature for carbon stars is the 11.3$\\mu$m emission due to the presence of SiC grains, which also represent an evidence of a C-rich dust envelope. Spitzer IRS spectra also confirmed two absorptions at 7 and 13.7 $\\mu$m, whose origin is still not well understood \\citep{speck06}. The first classification system for carbon-rich stars was the Henry Draper Catalogue \\citep{HD}, in which the stars were presented in two spectral classes: type R and N, that were also divided into temperature subclasses. These subclasses were based on lines in the blue spectral region, that, although being a good source of information to determine the temperature classes of G and K stars, were not an appropriate choice for the cold carbon stars. Therefore, the distinction between an R8, the redder R-type stars, from an Na, the most blue N-type, was not obvious. \\citet{MK41} decided to rearrange all the carbon-rich stars in another classification scheme that presented a single temperature sequence. The numerical temperature type was, at this time, determined based on the atomic and molecular structures, considered more susceptible to temperature variations. They also added a \\CC -index, based on the strength of the \\CC~bands. Many attempts were made to improve this MK system, while keeping its basic structure. \\citet{Yama72,Yama75} with the C-Classification System listed in his tables a very large number of stars with the two main parameters from the MK System, and additional intensity indices of several other atomic and molecular characteristics from carbon stars. The result of this attempt, was a detailed notation for carbon-rich objects, but it was not very practical or compact. Later, \\citet{Keenan93} pointed out several reasons for replacing this old C-Classification System, since much importance was given to the Na D-line in determining the numerical indices of temperature. The assumption that the Na D-line could be a good tracer of temperature in this case, however, was revealed to be quite unfortunate, specially in the case of N-type stars, as they have an enormous molecular opacity in the same spectral region due to the (7,2) CN band. Furthermore, the N-type and the R-type stars, in fact, describe two different populations that should not be classified under the same spectral class. The new classification system proposed by Keenan93, the MK~Revised System, re-established the spectral subclasses for the carbon rich objects and the temperature indices based on infrared intensities. Additionally, this Revised MK System listed four abundance indices from \\citet{Yama72}: the intensity of the \\CC~band, the isotopic carbon ratio, the Si\\CC\\ band and the CH band strength. These indices and system are the most widely accepted and used nowadays in the study of carbon stars \\citep{Morgan03,Morgan04,deroo07}. %%tentar achar uma referencia atual???? Concerning the circumstellar analysis, \\citet{sloan98} presented the first classification based on an infrared study of the spectra of carbon and oxygen-rich AGB stars. Qualitative classes for the circumstellar material were settled through two indices: one chemical and other a strength indicator of the characteristic structure of the emission spectrum. In the present paper we present a new scheme of classification for C-rich AGB stars (hereafter NSCC) that includes a notation conceived through a wide analysis of the main spectroscopic features of these carbon-rich objects with circumstellar material in its surroundings. The main difference between this scheme and the others available is that, instead of using a single region of the spectrum, we suggest a classification system based on a large wavelength range, from the blue-optical to the mid-infrared. Thus, this notation draws a more complete scenario of each star. The parameters used in this extended classification system were devised from the analysis of a sample of 51 observed AGB carbon stars. Section \\ref{obs} describes the optical observations and other infrared data used and section \\ref{optical} and \\ref{ir} details how indices and parameters of classification were obtained. While, section \\ref{optical} explains the optical indices, section \\ref{ir} contains the infrared ones. Section \\ref{disc} shows an analysis of the classification indices, a discussion of possible evolutionary sequences for the AGB carbon stars and also an atlas of optical spectra and infrared radiative transfer models of the stars studied in this work. The overall results are discussed in section \\ref{conc}. ", "conclusions": "\\label{conc} This New Scheme of Classification of C-Rich AGB Stars can not be applied to any carbon star. It is not a good application for AGB carbon stars that have a very thick envelope, e.g. extreme AGB stars can not be treated this way as their optical spectra are highly obscured. Moreover, the methodology employed is not tied to the sample presented, it has the flexibility to serve to all these kind of carbon stars by using the given coefficients and parameters. The indices correspond to either a direct measurement of the intensities and equivalent widths of features observed in low resolution spectra of carbon-rich stars or obtained through a radiative transfer model fit to infrared data. The seven indices presented describe in detail the complex scenario of the carbon rich stars. It is possible just by analyzing the compact final notation to get a full set of basic information about an AGB carbon star. As all calibrations were established based on the well quoted works, the indices and levels employed represent the more successful historical parameters on the study of carbon stars. Regarding the use of the indices, we demonstrated how the SiC/A.C. ratio and the $\\tau$ parameters could together provide information about the evolutionary stages of the carbon-rich stars. The evolutionary sequences here proposed may provide a new insight over the nature of objects such as the odd C-J stars. %% The \\notetoeditor{TEXT} command allows the author to communicate %% information to the copy editor. This information will appear as a %% footnote on the printed copy for the manuscript style file. Nothing will %% appear on the printed copy if the preprint or %% preprint2 style files are used. %###############################################" }, "0910/0910.2243_arXiv.txt": { "abstract": "We have developed a homogeneous model of physical chemistry to investigate the neutral-dominated, water-based Enceladus torus. Electrons are treated as the summation of two isotropic Maxwellian distributions$-$a thermal component and a hot component. The effects of electron impact, electron recombination, charge exchange, and photochemistry are included. The mass source is neutral H$_2$O, and a rigidly-corotating magnetosphere introduces energy via pickup of freshly-ionized neutrals. A small fraction of energy is also input by Coulomb collisions with a small population ($<$\\,1\\%) of supra-thermal electrons. Mass and energy are lost due to radial diffusion, escaping fast neutrals produced by charge exchange and recombination, and a small amount of radiative cooling. We explore a constrained parameter space spanned by water source rate, ion radial diffusion, hot-electron temperature, and hot-electron density. The key findings are: (1) radial transport must take longer than 12 days; (2) water is input at a rate of 100--180 kg s$^{-1}$; (3) hot electrons have energies between 100 and 250 eV; (4) neutrals dominate ions by a ratio of 40:1 and continue to dominate even when thermal electrons have temperatures as high as $\\approx$\\,5 eV; (5) hot electrons do not exceed 1\\% of the total electron population within the torus; (6) if hot electrons alone drive the observed longitudinal variation in thermal electron density, then they also drive a significant variation in ion composition. ", "introduction": "Absorption of UV starlight during occultation of Saturn's moon Enceladus showed that it continuously ejects neutral H$_2$O at a rate of $\\approx$\\,150--300\\,kg\\,s$^{-1}$ from water-ice geysers located at its southern pole \\citep{2006Sci...311.1422H}. Models suggest that the water and its chemical by-products form an extended neutral-dominated torus centered on the orbit of Enceladus \\citep{jurac2005}. Similarly, Jupiter's volcanic moon Io emits a mixture of SO$_2$ and S$_2$ at a rate of $\\approx$\\,1\\,ton\\,s$^{-1}$, and chemical by-products produce a plasma torus centered on the orbit of Io (see review by \\cite{thomas04}). The Hubble Space Telescope (HST) observations by \\cite{1993Natur.363..329S} revealed that the neutral-to-ion ratio in the Enceladus torus ($\\approx$\\,10) is three orders of magnitude greater than in the Io torus ($\\approx$\\,10$^{-2}$). Compositional differences and the degree of ionization within these two systems can be attributed to their chemistry (Io's based on sulfur dioxide and Enceladus's based on water) as well as the fact that fresh ions are picked with five times more energy in the Io torus than in the Enceladus torus \\citep{2007GeoRL..3409105D}. An important lesson learned from studying the physical chemistry of the Io torus is that a small fraction of hot electrons ($<$\\,1\\%) play a critical role in determining composition. To model the Cassini UltraViolet Imaging Spectrograph (UVIS) data obtained during Cassini's E2 flyby of Jupiter (October 2000 to March 2001), \\cite{steffl2,steffl1,steffl3,steffl4} adapted the \\cite{2003JGRA..108.1276D} Io torus model. \\textit{Steffl et al.}\\ used their models to study radial, temporal, and azimuthal variation in mixing ratios (ion-to-electron density ratios), thermal-electron density, and thermal-electron temperature. \\cite{steffl4} concluded that hot electrons are necessary for the Io torus energy budget and that two modulations of the hot-electron population are required to reproduce both the temporal and spatial variations in composition observed in the data, one modulating in Jupiter's System III longitude, the other in System IV. We anticipate that hot electrons are similarly important in Saturn's Enceladus torus. \\cite{2007GeoRL..3409105D} developed a simplified oxygen-based model to compare the Enceladus and Io tori. They found that collisional heating by a population of hot electrons is much less important at Enceladus, contributing only 0.5\\% of the energy to the torus, compared to 60\\% at Io. They also cited two major reasons for the discrepancy in the neutral-to-ion ratio between the two systems. First, newly created ions are picked up in the Io torus by Jupiter's magnetosphere at roughly five times the energies as are those in the Enceladus torus by Saturn's magnetosphere. The higher-energy pickup ions in the Io torus warm the thermal electrons, which then reduces the neutral-to-ion ratio via impact ionization. Second, because of the high abundance of molecular ions compared with atomic oxygen ions in the Enceladus torus (e.g., \\cite{sittler2005}), \\cite{2007GeoRL..3409105D} expected that molecular dissociative recombination (not included in their model) will therefore drive the ratio even higher in the Enceladus torus. From the conclusion of \\cite{2007GeoRL..3409105D}: ``The addition of the full water-group molecular chemistry will introduce an additional plasma sink through dissociative recombination of the molecular ions. Therefore, our simplified O-based chemistry likely represents a lower limit for the neutral/ion ratio.'' To investigate the consequences of a water-based Enceladus torus dominated by molecular chemistry, we have improved on the Delamere model by including a comprehensive set of water-based reactions and species to more accurately estimate steady-state densities and temperatures of ions and electrons. We also add neutral and ionized molecules. Molecules are more abundant in the Enceladus torus where low plasma density allows H$_2$O to escape from Enceladus largely intact, whereas the more energetic local plasma interaction at Io results in dissociation of SO$_2$ \\citep{2008JGRA..11309208D}. Moreover, thermal electrons ($\\approx$\\,2\\,eV) throughout the Enceladus torus dissociate H$_2$O approximately ten times less easily than thermal electrons ($\\approx$\\,5\\,eV) in the Io torus dissociate SO$_2$ (\\textit{V.\\ Dols, personal comm.}). Previous models of molecular chemistry in Saturn's magnetosphere were driven by Pioneer 11 and Voyager 1 and 2 observations \\citep{frank1980,trainor1980,wolfe1980,bridge1981,bridge1982,sittler1983}. \\cite{richardson1986} showed the importance of recombination under conditions of slow radial transport. Using essentially the same chemical reactions as \\cite{richardson1986}, \\cite{1998JGR...10320245R} determined ion and neutral lifetimes within Saturn's inner magnetosphere ($\\lesssim 12$$R_\\mathrm{S}$; $R_\\mathrm{S}\\equiv$ Saturn radius $ =6.0\\times10^9\\,\\mathrm{cm}$), constrained by HST observations of the extended OH cloud (see their Figure 2, and references therein). They solved the rate equations for number densities while also solving the radial diffusion equation, but energy conservation was not considered. \\cite{2002GeoRL..29x..25J} and \\cite{jurac2005} further improved on these models by considering neutral cloud expansion, and solved for plasma and neutral distributions self-consistently in order to study the source of water within Saturn's inner magnetosphere. Our model concentrates on the molecular chemistry. We start with a uniform box and characterize transport by just a time scale. However, we do consider energy balance. More importantly, unlike the above models, we retain H$_3$O$^+$ in our model, which proves to be a significant component. The purpose of this paper is to summarize the sensitivity of the chemistry of Enceladus's torus to several parameters. The parameters we investigated are hot-electron temperature, hot-electron density, H$_2$O source rate, ion radial-diffusion time scale, and proton temperature anisotropy. Being lighter, protons are less bound to the equator \\citep{1980GeoRL...7...41B,2008P&SS...56....3S,2009JGRA..11404211P}. This means protons spend only a fraction of the time interacting with the heavy ions and molecules. This effect is simulated with a `proton dilution' factor (Section \\ref{modl}). We search for values of these parameters leading to thermal-electron temperature, thermal-electron density, and water-group ($\\mathrm{W^ +\\equiv O^++OH^++H_2O^++H_3O^+}$) ion-to-proton ratio consistent with available Cassini data (Section \\ref{constraints}). The observations used to constrain our model and to define the parameter space are given in Section \\ref{data}. The model is described in Section \\ref{modl}. The best fit (baseline solution) and the procedure used to find it are discussed in Section \\ref{BaselineSolution}. Model sensitivity to each of the parameters is discussed in Section \\ref{grdSrch}. Finally, the importance of hot electrons with regard to water-group ion composition is demonstrated in Section \\ref{fehModulation}. ", "conclusions": "We have compared output from our model to the constraints on thermal-electron density ($n_\\mathrm{e}$), thermal-electron temperature ($T_\\mathrm{e}$), and mixing ratio of water-group ions to protons (W$^+/$H$^+$) by exploring the space spanned by the following four parameters: neutral source rate ($N_\\mathrm{src}$), hot-electron temperature ($T_\\mathrm{eh}$), hot-electron density ($n_\\mathrm{eh}\\equiv f_\\mathrm{eh}n_\\mathrm{e}$), and radial transport time scale ($\\tau_\\mathrm{trans}$). Our important results are: \\begin{enumerate} \\item For the constraint choices of $n_\\mathrm{e}=60\\ \\mathrm{cm}^{-3}$, $T_\\mathrm{e}=2$ eV, and W$^+$/H$^+$=12, we find the following limits on the parameters: \\begin{enumerate} \\item[\\ ] $1.5\\lesssim N_\\mathrm{src}/(10^{-4}\\mathrm{\\ cm^{-3}\\ s^{-1}})\\lesssim 2.7$ \\item[\\ ] $12\\ \\mathrm{days} \\lesssim \\tau_\\mathrm{trans}$ \\item[\\ ] $0.3\\lesssim f_\\mathrm{eh}(\\%)\\lesssim0.9$ \\item[\\ ] $100\\lesssim T_\\mathrm{eh}/\\mathrm{eV}\\lesssim250$. \\end{enumerate} \\noindent For a volume of $2\\pi(4R_\\mathrm{S})(2R_\\mathrm{S})^2$, the source rate can be scaled to give a mass source rate of $100\\lesssim N_\\mathrm{src}/(\\mathrm{kg\\ s^{-1}})\\lesssim 180$. We find that $f_\\mathrm{H^+}$ (the fraction of protons confined to the equator) is strongly coupled to the hot-electron population and has not been constrained by this study (Section \\ref{varyingProtonDilution}). The solution space can be compared with future measurements of the parameters ($N_\\mathrm{src},\\ f_\\mathrm{eh},\\ T_\\mathrm{eh},\\ \\tau_\\mathrm{trans}$) and composition mixing ratios. Upper limits on UV power emanating from the Enceladus torus from neutral oxygen at 1304, 1356, and 6300\\,\\AA\\ would be very useful. \\item With the full water-based chemistry, photo- plus impact ionization is nearly as important as charge exchange at providing energy by way of fresh pickup ions (Figure \\ref{enerFlow}). \\item The water-based chemistry (particularly dissociative recombination) increases the neutral-to-ion ratio from 12 (the \\cite{2007GeoRL..3409105D} oxygen-based model) to $\\approx 40$. Further, the Enceladus torus remains neutral-dominated even when the thermal-electron temperature approaches the temperature of electrons in Jupiter's Io torus (6 eV). \\item The H$_3$O$^+$/W$^+$ ratio is directly correlated with the W$^+$/H$^+$ ratio (Figure \\ref{barchart}), implying that H$_3$O$^+$ is strongly anti-correlated with H$^+$ (Section \\ref{alternativeConstraints}). This result is important because obtaining the W$^+$/H$^+$ ratio is more straightforward than obtaining individual water-group abundances from the CAPS data. However, \\cite{2008P&SS...56....3S} have shown, based on CAPS SOI data, that H$_3$O$^+$ dominates the water-group near the orbit of Enceladus (their Figure 15). The dominance of H$_3$O$^+$ seen in the CAPS data has not been obtained by our model with the given constraints. It is likely (manuscript in preparation) that the local interaction of the corotating plasma with the Enceladus plumes may contribute significantly to the H$_3$O$^+$ abundance \\citep{cravens2009}. \\item Hot electrons are necessary in the Enceladus to enhance ionization torus but do not directly contribute more that 1\\% of the total electron population. \\item Significant variation in composition can be driven by a small perturbation in the hot-electron population (Figure \\ref{fehVoutput}). \\end{enumerate} The sensitivity study presented here will be useful in a future interpretation of longitudinal and temporal observations (e.g., \\cite{gurnett2007}) of the Enceladus torus, especially in the context of a modulating hot-electron density." }, "0910/0910.5180_arXiv.txt": { "abstract": "Galaxies, in particular disc galaxies, contain a number of structures and substructures with well defined morphological, photometric and kinematic properties. Considerable theoretical effort has been put into explaining their formation and evolution, both analytically and with numerical simulations. In some theories, structures form during the natural evolution of the galaxy, i.e. they are a result of nature. For others, it is the interaction with other galaxies, or with the intergalactic medium -- i.e. nurture -- that accounts for a structure. Either way, the existence and properties of these structures reveal important information on the underlying potential of the galaxy, i.e. on the amount and distribution of matter -- including the dark matter -- in it, and on the evolutionary history of the galaxy. Here, I will briefly review the various formation scenarios and the respective role of nature and nurture in the formation, evolution and properties of the main structures and substructures. ", "introduction": "What do we mean by ``nature versus nurture''? To some degree, all galaxies have one or more neighbours and/or are influenced by their surroundings. So in the absolute sense of the word, there is no such thing as an isolated galaxy, except in computer simulations. On the other hand, even though a given structure may be triggered by an interaction, it will evolve within its galaxy potential, which, together with the remaining characteristic properties of the galaxy, will influence its evolution. Thus, no structure is 100\\% due to nature, or 100\\% due to nurture. It is, nevertheless, useful to ask which of the two has mainly influenced the formation and evolution of a given structure, and, therefore, in the following I will assign the origin of a structure to one of these two alternatives. In fact, as will be seen below, more than one scenario is possible for each structure, and, very often, some mixture of nature and nurture can be involved. Hence, one can only reason in terms of probabilities and propose this or that agent as the most probable cause for a given feature. Time comes into play as well: if one goes sufficiently far back, all galaxies have experienced important interactions with their surroundings. If, however, the time elapsed since the last interaction is longer than the time necessary for a given structure to form, one can safely argue that this structure is due to nature rather than to nurture (see also a discussion in \\cite{HopkinsKMQ09b}). After these introductory remarks, let me review what models and simulations can tell us. Since the subject is very broad, I will discuss here only certain aspects, necessarily reflecting my own preferences, without me wanting to belittle in any way work which I cannot mention for lack of time and space. The list of references given is likewise far from complete. Finally, I will only discuss the formation of structures and not their destruction, although the latter could also be due either to nature or to nurture. ", "conclusions": "I discussed the possible formation mechanisms of structures and substructures in galaxies, focusing on whether their origin is due to nature or to nurture. In most cases, there is more than one alternative. A few structures can be explained only by a mechanism relying on nurture, but none uniquely only to nature." }, "0910/0910.1595_arXiv.txt": { "abstract": "We have undertaken a deep ($\\sigma\\sim 1.1$ mJy) 1.1-mm survey of the $z=0.54$ cluster MS\\,0451.6$-$0305 using the AzTEC camera on the James Clerk Maxwell Telescope. We detect 36 sources with S/N$\\ge 3.5$ in the central 0.10 deg$^2$ and present the AzTEC map, catalogue and number counts. We identify counterparts to 18 sources (50\\%) using radio, mid-infrared, \\Spitzer\\ IRAC and Submillimeter Array data. Optical, near- and mid-infrared spectral energy distributions are compiled for the 14 of these galaxies with detectable counterparts, which are expected to contain all likely cluster members. We then use photometric redshifts and colour selection to separate background galaxies from potential cluster members and test the reliability of this technique using archival observations of submillimetre galaxies. We find two potential MS\\,0451$-$03 members, which, if they are both cluster galaxies have a total star-formation rate (SFR) of $\\sim100$ \\myr\\ -- a significant fraction of the combined SFR of all the other galaxies in MS\\,0451$-$03. We also examine the stacked rest-frame mid-infrared, millimetre and radio emission of cluster members below our AzTEC detection limit and find that the SFRs of mid-IR selected galaxies in the cluster and redshift-matched field populations are comparable. In contrast, the average SFR of the morphologically classified late-type cluster population is $\\sim 3$ times less than the corresponding redshift-matched field galaxies. This suggests that these galaxies may be in the process of being transformed on the red-sequence by the cluster environment. Our survey demonstrates that although the environment of \\ms\\ appears to suppress star-formation in late-type galaxies, it can support active, dust-obscured mid-IR galaxies and potentially millimetre-detected LIRGs. ", "introduction": "\\label{sec:intro} Galaxy clusters are highly biased environments in which galaxies potentially evolve more rapidly than in the field. The galaxy populations of local massive clusters contain mainly early-type galaxies which define a colour-magnitude relation (CMR) \\citep{Visvanathan77, Bower92}. However, studies of clusters out to $z\\sim 1$ suggest that they contain increasing activity at higher redshifts due to a growing fraction of blue, star-forming galaxies \\citep{Butcher84}. Over the same redshift range there appears to be a growing deficit in the CMR population at faint magnitudes, as well as a claimed increasing decline in the numbers of S0 galaxies, suggesting that the blue, star-forming galaxies may be transforming into these passive populations with time \\citep{Dressler97, Smail98a, DeLucia07, Stott07, Holden09}. The blue, star-forming populations within the clusters are accreted from the surrounding field as the clusters grow via gravitational collapse. The evolution in the star-forming fraction in the clusters may thus simply track the increasing activity in the field population at higher redshifts. The increasing activity in the field is also reflected in an increasing number of the most luminous (and dusty) star-burst galaxies with redshift \\citep[e.g.][]{LeFloch05}: the Luminous Infrared Galaxies (with L$_{\\rm FIR}\\geq 10^{11}$\\,L$_\\odot$) and their Ultraluminous cousins (ULIRGs, L$_{\\rm FIR}\\geq 10^{12}$\\,L$_\\odot$). These systems will also be accreted into the cluster environment along with their less-obscured and less-active population as the clusters grow. Indeed, mid-infrared (mid-IR) surveys of clusters have identified a population of dusty star-bursts whose activity increases with redshift \\citep[e.g.][]{Geach06, Geach08}. However, these mid-IR surveys can miss the most obscured (and potentially most active) galaxies which are optically thick in the rest-frame mid-IR. If they are present in clusters -- even in relatively low numbers -- such active galaxies will contribute significantly to the star formation rate (SFR) in these environments and the metal enrichment of the intracluster medium. Hence to obtain a complete census of the star formation within clusters we need to survey these systems at even longer wavelengths, in the rest-frame far-infrared (far-IR), corresponding to the observed sub-millimetre and millimetre waveband. Over the past decade or more there have been a number of studies of rich clusters of galaxies in the sub-millimetre and millimetre wavebands \\citep[e.g.][]{Smail97, Zemcov07, Knudsen08}. Many of these studies were seeking to exploit the clusters as gravitational telescopes to aid in the study of the distant sub-millimetre galaxy (SMG) population behind the clusters, while others focused on the detection of the Sunyaev-Zel'dovich (SZ) emission. Due to the limitations of current technologies direct detection of millimetre sources is restricted to those with fluxes S$_{\\rm 1100\\umu m}\\ga 1$\\,mJy, or equivalently galaxies with SFRs $\\ga 300$\\,M$_{\\sun}$\\,yr$^{-1}$ -- much higher than expected for the general cluster populations based on optical surveys. Nevertheless these studies have serendipitously detected a number of cluster galaxies, although these are either atypical (e.g.\\ central cluster galaxies, \\citet{Edge99}) or are not confirmed members \\citep[e.g.][]{Best02, Webb05}. More critically, with few exceptions these studies have all focused on the central 2--3 arcmin of the clusters, where the SZ emission and lensing amplification both peak, and have not surveyed the wider environment of the cluster outskirts where much of the obscured star formation is likely to be occurring \\citep[e.g.][]{Geach06}. The two exceptions are the wide-field survey of the $z\\sim0.25$ cluster A\\,2125 by \\citet{Wagg09} and the survey of an overdense region of the COSMOS field by \\citet{Scott08} and \\citet{Austermann09a}. \\citet{Wagg09} detected 29 millimetre galaxies across a $\\sim 25$-arcmin diameter region centered on A\\,2125 of which none are claimed to be members. However, the only redshift estimates available are based on crude radio-to-millimetre spectral indices, which are sensitive to both dust temperature and redshift \\citep{Blain03}. The AzTEC/COSMOS survey \\citep{Scott08, Austermann09a} covered a number of structures, including an X-ray detected $z=0.73$ cluster and concluded that the statistical overdensity of sources was most likely due to the gravitational lensing of background SMGs by these foreground structures. To help to conclusively answer the issue of the obscured star-forming populations in distant clusters we have therefore undertaken a wide-field 1.1-mm survey of the $z=0.54$ rich cluster MS\\,0451.6$-$0305 (hereafter \\ms) with the AzTEC camera \\citep{Wilson08} on the James Clerk Maxwell Telescope (JCMT). This panoramic millimetre survey can also take advantage of the significant archival data available for this well-studied region. In particular the panoramic imaging of \\ms\\ from {\\it Spitzer} and uniquely, the {\\it Hubble Space Telescope} ({\\it HST}), as well as extensive archival multiwavelength imaging and spectroscopy, which we employ for determining cluster membership of AzTEC-detected galaxies. Our survey probes the rest-frame 700\\,$\\umu$m emission of cluster galaxies -- in search of examples of strongly star-forming but obscured galaxies -- as well as identifying more luminous and more distant examples of the SMG population. We can compare our millimetre search for cluster members to the previous mid-IR survey of this cluster by \\citet{Geach06} who uncovered a population of dusty star-forming galaxies which dominate the integrated SFR of the cluster of $200\\pm 100$\\,M$_{\\sun}$\\,yr$^{-1}$. Our AzTEC map covers $\\sim 60$ times the area of the 850\\,$\\umu$m SCUBA observations of the central region of \\ms\\ \\citep{Borys04a}, while the depth of $\\sigma\\sim1.1$ mJy is sufficient to identify ULIRGs individually and obtain stacked constraints on far-IR fainter cluster members. We describe our observations and the data reduction in \\S\\ref{sec:obsred}; present 1.1-mm number counts, identify counterparts to the AzTEC galaxies, and use photometric techniques to determine cluster membership in \\S\\ref{sec:analysis}. \\S\\ref{sec:conc} contains our conclusions; individual AzTEC galaxies are presented in Appendix \\ref{sourcenotes}. Throughout this paper we use J2000 coordinates and $\\Lambda$CDM cosmology with $\\Omega_M=0.3$, $\\Omega_{\\Lambda}=0.7$ and $H_0=70$ ${\\rm kms^{-1}Mpc^{-1}}$ and all photometry is on the AB system unless otherwise stated. ", "conclusions": "\\label{sec:conc} In this study we have investigated the dust-obscured star-forming population of the galaxy cluster \\ms; we utilise 1.1-mm observations to study obscured star-formation in \\ms. We present a $\\sigma\\sim 1.1$ mJy AzTEC map of the central 0.1 deg$^2$ of \\ms, within which 36 sources are detected at S/N$\\ge3.5$. We use radio, 24\\,$\\umu$m, IRAC and SMA observations to precisely locate 18 of these SMGs. We calculate the reliability of photometric redshifts for SMGs and find that they are able to remove the bulk of background contamination, allowing us to isolate potential cluster members using our optical, near- and mid-IR photometry. Based on these redshifts we find two SMGs which are possible cluster members: MMJ\\,045421.55 and MMJ\\,045431.35. These systems are both resolved into close pairs of galaxies by our ground-based and \\HST\\ imaging, suggesting that interactions have triggered their starbursts. If they are cluster members both of these SMGs contain cold dust with $T_d\\sim30$ K, for $\\beta=1.1$ \\citep[similar to SLUGS galaxies;][]{Dunne00a} and have SFRs of $\\sim50$ \\myr\\ each. \\citet{Geach06} compared obscured activity, based on $24 \\umu$m emission, in \\ms\\ and \\cl\\ at $z=0.39$, and found that \\cl\\ has ${\\rm SFR}\\sim5$ times that of \\ms\\ within 2 Mpc of the clusters centres. They show that, taking into account the slightly different redshifts and masses of these two structures, \\ms\\ is underactive and \\cl\\ overactive at $24 \\umu$m. Therefore, it is likely the other galaxy clusters, including \\cl, could contain ULIRGs in significantly larger numbers than we find in \\ms. To further investigate the obscured star-forming population which lies below the limit of our AzTEC observations we create composite SEDs of spectroscopically confirmed mid-IR, early- and late-type cluster members and compare them to the corresponding field populations. As expected we find that both early-type populations are undetected at both 24\\,$\\umu$m and 1.1 mm and so are unlikely to be actively forming large numbers of stars ($<0.1$ \\myr). The 24\\,$\\umu$m galaxies are significantly more active than the morphologically classified late-types with SFRs $\\sim15$ \\myr\\ versus $\\sim3$ \\myr\\ on average. We find that the star-formation activity in the cluster late-type population, compared to a redshift-matched field population is quenched and $\\sim3$ times lower than expected. Mid-IR galaxies do not show this trend suggesting the more intense activity in these systems is more robust to environmental influences. We find that the total ${\\rm SFR>315\\pm50}$ \\myr\\ in the central 2 Mpc of \\ms. However, if MMJ\\,045421.55 is a cluster member it has ${\\rm SFR}\\sim50$ \\myr\\ and lies 1 Mpc from the cluster centre, taking the total SFR within 2 Mpc to $\\ga360$ \\myr." }, "0910/0910.0962_arXiv.txt": { "abstract": "In the context of coronal heating, among the zoo of MHD waves that exist in the solar atmosphere, Alfv\\'en waves receive special attention. Indeed, these waves constitute an attractive heating agent due to their ability to carry over the many different layers of the solar atmosphere sufficient energy to heat and maintain a corona. However, due to their incompressible nature these waves need a mechanism such as mode conversion (leading to shock heating), phase mixing, resonant absorption or turbulent cascade in order to heat the plasma. Furthermore, their incompressibility makes their detection in the solar atmosphere very difficult. New observations with polarimetric, spectroscopic and imaging instruments such as those on board of the japanese satellite \\textit{Hinode}, or the \\textit{SST} or \\textit{CoMP}, are bringing strong evidence for the existence of energetic Alfv\\'en waves in the solar corona. In order to assess the role of Alfv\\'en waves in coronal heating, in this work we model a magnetic flux tube being subject to Alfv\\'en wave heating through the mode conversion mechanism. Using a 1.5-dimensional MHD code we carry out a parameter survey varying the magnetic flux tube geometry (length and expansion), the photospheric magnetic field, the photospheric velocity amplitudes and the nature of the waves (monochromatic or white noise spectrum). The regimes under which Alfv\\'en wave heating produces hot and stable coronae is found to be rather narrow. Independently of the photospheric wave amplitude and magnetic field a corona can be produced and maintained only for long ($> 80$ Mm) and thick (area ratio between photosphere and corona $> 500$) loops. Above a critical value of the photospheric velocity amplitude (generally a few km s$^{-1}$) the corona can no longer be maintained over extended periods of time and collapses due to the large momentum of the waves. These results establish several constraints on Alfv\\'en wave heating as a coronal heating mechanism, especially for active region loops. ", "introduction": "New observations from polarimetric, spectroscopic and imaging instruments are revealing a corona permeated with waves. Compressive modes such as the slow and fast magnetohydrodynamic (MHD) modes have a long chapter in observational history \\citep[see reviews by ][]{Nakariakov_2005LRSP....2....3N, Banerjee_2007SoPh..246....3B, Ruderman_2009SSRv..tmp...54R, Taroyan_2009SSRv..tmp...24T}. Only recently, the development of high resolution instruments has brought strong evidence for the existence of the third MHD mode, the Alfv\\'en mode, in the solar atmosphere. In the majority of the observational reports of Alfv\\'en waves ambiguity exists and the waves can be interpreted as trapped modes, fast and slow kink waves. \\citet{DePontieu_2007Sci...318.1574D} reported transversal displacements of spicules from observations in the Ca II H-line (3968 \\AA) with a broadband filter of \\textit{Hinode}/SOT. The wavelengths of these chromospheric agents were estimated to be longer than 20000 km, with periods between 100 and 500 s and speeds of at least 50-200 km s$^{-1}$. In the absence of a stable waveguide they interpreted the swaying of the spicules as a result of upward propagating Alfv\\'en waves. However, it was pointed out by \\citet{Erdelyi_2007Sci...318.1572E} that these waves were likely to be kink oscillations, due to the displacement of the axis of symmetry of the flux tube caused by the later. Using the Coronal Multi-Channel Polarimeter (\\textit{CoMP}) \\citet{Tomczyk_2007Sci...317.1192T} analyzed properties of infrared coronal emission line Fe\\,XIII (1074.7 nm) across a large field of view with short integration times. Waves propagating along magnetic field lines with ubiquitous quasi-periodic fluctuations in velocity with periods between 200 and 400 s (power peak at 5-min) and negligible intensity variations were reported. Wavelengths and phase speeds were estimated to be higher to 250 Mm and 1 Mm s$^{-1}$ respectively. In this work it was suggested that these waves were Alfv\\'en waves probably being generated in the chromospheric network from mode conversion of p-modes propagating from the photosphere (hence explaining the peak in the power spectrum of the waves). On the other hand \\citet{Erdelyi_2007Sci...318.1572E} and \\citet{VanDoorsselaere_2008ApJ...676L..73V} argued that an interpretation in terms of fast kink waves was more appropriate due mainly to the observed collective behavior which should be absent in the case of Alfv\\'en waves. Being compressible waves it was argued that kink waves would appear however incompressible in the corona for an instrument such as \\textit{CoMP} due to the very small variation in intensity they produce. The difficulty in detecting the Alfv\\'en wave is due in part to its incompressible nature. Being only a transverse disturbance in the magnetic field, it is practically invisible to imaging instruments, unless the structure of propagation is displaced periodically from the line of sight \\citep{Williams_2004ESASP.547..513W}. Also, due to the large wavelengths (a few megameters) and short periods (a few minutes) that may be involved in the corona, a large field of view, short integration time and proper resolution are needed for their detection in the corona. With polarimeters and spectrographs however Alfv\\'en waves are more easily detected, as suggested by \\citet{Erdelyi_2007Sci...318.1572E}. These waves propagate as torsional disturbances of the magnetic field, causing a periodic and spatially dependent spectral line broadening \\citep{Zaqarashvili_2003AA...399L..15Z}. Up to date, the only observational report on torsional Alfv\\'en waves is the recent work by \\citet{Jess_2009Sci...323.1582J}, in which observations in H$\\alpha$ of a bright point using the Solar Optical Universal Polarimeter (SOUP) of the \\textit{SST} were reported. No variation in intensity nor line-of-sight velocity was detected and a periodic spectral broadening of the H$\\alpha$ line was found, leaving Alfv\\'en waves as the only possible interpretation for their results. The waves were reported to have periods in the 100-500 s interval (with maximum power in the 400-500 s range). By comparing H$\\alpha$ continuum with H$\\alpha$ core images they reported a canopy-like structure with magnetic fields expanding $\\sim1300$ km in a vertical distance of $\\sim1000$ km. This large expansion is crucial for Alfv\\'en wave heating theories. For instance, assuming an Alv\\'en speed of 10 km s$^{-1}$ in the upper photosphere/ lower chromosphere and a period of 100 s (a value in the lower range of periods reported in the works cited above) we obtain a wavelength of 1000 km for Alfv\\'en waves propagating from the photosphere, matching the distance of large expansion of the flux tubes. Moreover, the height in the atmosphere where this takes place is the region of transition from high beta to low beta plasma. We thus have ideal conditions for mode conversion. Trapped modes will thus exchange their energy and a wide range of waves with different characteristics are likely to issue from this region. The large emphasis that has been put on the search for Alfv\\'en waves in the solar corona is largely due to their connection with coronal heating. Theoretically, they can be easily generated in the photosphere by the constant turbulent convective motions, which inputs large amounts of energy into the waves \\citep{Muller_1994AA...283..232M, Choudhuri_1993SoPh..143...49C}. Having magnetic tension as its restoring force the Alfv\\'en wave is less affected by the large transition region gradients with respect to other modes. Also, when traveling along thin magnetic flux tubes they are cut-off free since they are not coupled to gravity (Musielak et al. 2007)\\footnote{\\citet{Verth_etal_aap_09} have pointed out however that the assertion made by \\citet{Musielak_2007ApJ...659..650M} is valid only when the temperature in the flux tube does not differ from that of the external plasma. When this is not the case a cut-off frequency is introduced.}. Alfv\\'en waves generated in the photosphere are thus able to carry sufficient energy into the corona to compensate the losses due to radiation and conduction, and, if given a suitable dissipation mechanism, heat the plasma to the high million degree coronal temperatures \\citep{Uchida_1974SoPh...35..451U, Wentzel_1974SoPh...39..129W, Hollweg_1982SoPh...75...35H, Poedts_1989SoPh..123...83P, Ruderman_1997AA...320..305R, Kudoh_1999ApJ...514..493K} and power the solar wind \\citep{Suzuki_2006JGRA..11106101S, Cranmer_2007ApJS..171..520C}. Another possible generation mechanism for Alfv\\'en waves is through magnetic reconnection. The amount of energy imparted to these waves during the reconnection process may depend on the location in the atmosphere of the event and is a subject of controversy. \\citet{Parker_1991ApJ...372..719P} suggested a model in which 20~\\% of the energy released by reconnection events in the solar corona is transfered as a form of Alfv\\'en wave. \\citet{Yokoyama_1998ESASP.421..215Y} studied the problem simulating reconnection in the corona, and found that less than 10~\\% of the total released energy goes into Alfv\\'en waves. This result is similar to the 2-D simulation results of photospheric reconnection by \\citet{Takeuchi_2001ApJ...546L..73T}, in which it is shown that the energy flux carried by the slow magnetoacoustic waves is one order of magnitude higher that the energy flux carried by Alfv\\'en waves. On the other hand, recent simulations by Kigure et al. (private communication) show that the fraction of Alfv\\'en wave energy flux in the total released magnetic energy during reconnection in low $\\beta$ plasmas may be significant (more than 50~\\%). Since the observed ubiquitous intensity bursts (nanoflares) are thought to play an important role in the heating of the corona \\citep{Hudson_1991SoPh..133..357H} and since they are generally assumed to be a signature of magnetic reconnection it is then crucial to determine the energy going into the Alfv\\'en waves during the reconnection process. Moreover, it is possible that the observed nanoflares themselves are a consequence of Alfv\\'en wave heating as porposed by\\citet{Moriyasu_2004ApJ...601L.107M}. In that work the resulting intensity bursts producing the nanoflares are created by Alfv\\'en waves by first converting to longitudinal modes which steepen into shocks and heat the plasma. This model was further developed by \\citet{Antolin_2008ApJ...688..669A} (hereafter, Paper 1) and \\citet{Antolin_2009arXiv0903.1766A}, where it was shown that the frequency of the resulting heating events from Alfv\\'en waves followed a power law distribution. It was further shown that Alfv\\'en wave heating could be differentiated from nanoflare-reconnection heating during observations through a series of signatures \\citep[see also][]{Taroyan_2008IAUS..247..184T, Taroyan_2009SSRv..tmp...24T}. For instance, Alfv\\'en waves lead to a dynamic, uniformly heated corona with steep power law indexes (issuing from statistics of heating events) while nanoflare-reconnection heating leads to lower dynamics (unless catastrophic cooling takes place in the case of footpoint concentrated heating) and shallow power laws. The main problem faced by Alfv\\'en wave heating is to find a suitable dissipation mechanism. Being an incompressible wave it must rely on a mechanism by which to convert the magnetic energy into heat. Several dissipation mechanisms have been proposed, such as parametric decay \\citep{Goldstein_1978ApJ...219..700G, Terasawa_1986JGR....91.4171T}, mode conversion \\citep{Hollweg_1982SoPh...75...35H, Kudoh_1999ApJ...514..493K, Moriyasu_2004ApJ...601L.107M}, phase mixing \\citep{Heyvaerts_1983AA...117..220H, Ofman_2002ApJ...576L.153O}, or resonant absorption \\citep{Ionson_1978ApJ...226..650I, Hollweg_1984ApJ...277..392H, Poedts_1989SoPh..123...83P, Erdelyi_1995AA...294..575E}. The main difficulty lies in dissipating sufficient amounts of energy in the correct time and space scales. For more discussion regarding this issue the reader can consult for instance \\citet{Klimchuk_2006SoPh..234...41K, Erdelyi_2007AN....328..726E} and \\citet{Aschwanden_2004psci.book.....A}. Due to the observed increasing importance of Alfv\\'en waves in the solar atmosphere in this work we address the subject of coronal heating by Alfv\\'en waves in which mode conversion and parametric decay are taken as dissipation mechanisms. We concentrate on the heating of closed magnetic structures which populate the solar atmosphere, such as coronal loops, and analyze the efficiency of this heating model by carrying out a parametric space survey. We take a 1.5-dimensional model and we consider different photospheric drivers (a random driver creating a white noise spectrum, or a monochromatic driver in which several periods are tested), vary the photospheric magnetic field, the loop expansion from the photosphere to the corona and the loop length, and determine the regimes for which Alfv\\'en wave heating plays an important role in coronal heating. The paper is organized as follows. In \\S\\ref{two} we set up the loop geometries and define the equations. As in Antolin et al. (2008) the loops are modeled with a 1.5-dimensional MHD code including thermal conduction and radiative cooling. In \\S\\ref{three} we analyze the effect of varying the parameters on the thermodynamic structure of the loop. Further discussion and conclusions are presented in \\S\\ref{four}. ", "conclusions": "\\label{four} Alfv\\'en waves constitute one of the main candidates for heating the solar corona. Theoretically easy to be generated from convective turbulent motions in the photosphere it has been shown that these waves are able to carry sufficient amounts of energy in order to balance energy losses from radiation and conduction and heat the corona up to the observed few million degrees \\citep{Uchida_1974SoPh...35..451U, Wentzel_1974SoPh...39..129W, Hollweg_1982SoPh...75...35H, Poedts_1989SoPh..123...83P, Ruderman_1997AA...320..305R, Kudoh_1999ApJ...514..493K}. Finding a suitable dissipation mechanism is a very difficult task, which has spawned a large active research community in solar physics. In this work we have considered a model in which mode conversion acts as the dissipative mechanism. More precisely, through nonlinear effects based on density fluctuations and wave-to-wave interaction in the chromosphere and corona the Alfv\\'en mode is able to transfer some part of its energy to the fast and slow modes, which steepen into shocks and heat the plasma. We have conducted a parameter survey in which the effect of geometrical quantities such as the loop expansion and the loop's length on the efficiency of Alfv\\'en wave heating is studied. We further investigate the effect on the heating of other physical parameters such as the photospheric magnetic field and the generation of Alfv\\'en waves with a monochromatic and white noise spectrum in the photosphere. In Fig.~\\ref{fig8} we plot the volumetric heating rate with respect to the considered parameters of the loop: period of the monochromatic longitudinal waves issuing from mode conversion of Alfv\\'en waves (top left panel), photospheric magnetic field (top right panel), expansion of the loop (bottom left panel), and loop length (bottom right panel). The volumetric heating rate for each case corresponds to its extremum values in the range between $\\langle v_{\\phi}^{2}\\rangle^{1/2}\\sim1.3$ and 2.3 km s$^{-1}$ found in the e panels of Figs.~\\ref{fig3}, \\ref{fig5}, \\ref{fig6} and \\ref{fig7}, which is the most efficient amplitude range for coronal heating we have found. We can see that the volumetric heating rate is an increasing function for all the considered parameters in the chosen azimuthal photospheric velocity interval. This figure summarizes well the results of our parameter survey. As the amplitudes of the twists in the photosphere generating the Alfv\\'en waves increase the momentum and energy flux of the waves increase. We have found that nonlinear effects generally increase as well, and mode conversion happens not only in the chromosphere but ubiquitously in the corona, thus heating the plasma. Suitable conditions for Alfv\\'en wave heating are thus strongly dependent on the nonlinear effects. The importance over nonlinearity favors thick and long loops (bottom left and right panels), and strong photospheric magnetic fields (top right panel). The dependence on the later is however not crucial and flux tubes with 1 kG field concentrations at their footpoints may be heated efficiently by Alfv\\'en waves. On the other hand, the expansion and the length of the loop are crucial parameters for the heating. As we can see in Fig.~\\ref{fig8}, the volumetric heating rate in the bottom left and right panels does not vary much respect to the loop expansion and the length, which shows the high sensitivity of the thermodynamics in the loop on the heating. Nonlinearity is directly proportional to the loop expansion. Furthermore, the strong expansion of loops from the photosphere to the transition region happens in a height comparable to the wavelength of the Alfv\\'en wave, which leads to the deformation of the Alfv\\'en wave and excitation of slow modes \\citep{Moore_1991ApJ...378..347M}. Since the plasma $\\beta$ parameter in that region is close to unity, further mode conversion is expected. Alfv\\'en waves have long wavelengths and in order for them to convert into longitudinal waves which steep into shocks they need to propagate relatively long distances. Loops with lengths lower than 80 Mm cannot be heated by Alfv\\'en waves in the present model. Caution must be taken however when drawing conclusions from these results. When calculating the lengths of loops in observations what is actually calculated is the length of the corona, which is the visible part of loops in EUV or X-ray images. In our model, the lengths of the coronae with average temperatures of $\\sim1$ MK we obtain with Alfv\\'en wave heating oscillate between 50 and 80 Mm with a mean around 70 Mm. Furthermore, we have defined a stable corona, as a corona which can be maintained over roughly 5.7 hours. However, observations show that coronal loops are dynamical entities which exhibit heating and cooling processes constantly. Likewise, the loops with a collapsing corona obtained here exhibit heating events with nanoflare-like temperatures constantly, which can match the observed bursty X-ray intensity profiles and the statistics displaying power law distribution of the heating events (cf. Paper 1). Also, a loop which is visible in soft X-rays over a large period of time may in fact result from the effect of different threads in the loop being heated at different times, thus giving the impression of a steady and uniformly heated loop \\citep[as discussed in][]{Patsourakos_2009ApJ...696..760P}. In the simple model we have considered we have assumed that the heating acts uniformly along the radial direction, thus disabling the possibility of many threads being heated at different times. We thus cannot completely reject the possibility of Alfv\\'en waves heating the coronae of short and thin loops. This idea should however be tested in an extended (at least 2.5 dimensional) version of this model. All coronae show a range of velocities for which they are stable throughout the simulation and for which their lengths are constant. However, when the photospheric velocity field exceeds a critical value the corona collapses. The critical velocity is dependent on the parameters of the model. In this regime wave heating can no longer account for the large cool mass flux from the increasing wave amplitudes and collapses. Generally, the collapse is not due to a local increase in the corona of the radiative losses from the formation of a cool condensation as in the case of catastrophic cooling \\citep{Antiochos_1999ApJ...512..985A}, and does not exhibit ``limit cycles\" \\citep{Muller_2003AA...411..605M}. The later are characteristic of the catastrophic cooling mechanism which is proposed as an explanation for coronal rain. Further investigation in that direction is however required and will be the subject of a future paper. All of the obtained coronae are uniformly heated irrespective of the parameters in the model. This is a characteristic of Alfv\\'en wave heating. The distribution of the heating in coronal loops as derived from observations is a matter of debate. Indeed, different methods for the analysis of observational data may lead to different results. A famous example of this fact is the analysis of the data set from soft X-ray observations with the \\textit{Yohkoh}/SXT instrument. An interpretation of the heating distribution of the observed loops in terms of apex concentrated heating \\citep{Reale_2002ApJ...580..566R}, footpoint concentrated heating \\citep{Aschwanden_2001ApJ...559L.171A} and uniform heating \\citep{Priest_1998Natur.393..545P} has been given. There seems to be however more observational evidence of coronal loops being heated towards the footpoints in active regions \\citep{Aschwanden_2001ApJ...560.1035A, Aschwanden_2000ApJ...541.1059A}. Further evidence of this fact has been found by \\citet{Hara_2008ApJ...678L..67H} using the \\textit{Hinode}/EIS instrument, in which active region loops are shown to exhibit upflow motions and enhanced nonthermal velocities in the hot lines of Fe XIV 274 and Fe XV 284. Possible unresolved high-speed upflows were also found. In Paper 1 we found that footpoint or uniform nanoflare heating exhibit hot upflows, thus fitting in the observational scenario of active regions, while Alfv\\'en wave heating was found to exhibit hot downflows, which may fit in the observational scenario of quiet Sun regions \\citep{Chae_1998ApJS..114..151C, Brosius_2007ApJ...656L..41B}. The uniform heating from Alfv\\'en waves further supports this conclusion, since it is a characteristic of loops more generally found in quiet Sun regions. Active region loops exhibit low expansion factors due to the high magnetic field filling factor in those regions. In this chapter we have found that Alfv\\'en wave heating is effective only in thick loops (with area expansions between photosphere and corona higher than 600), further emphasizing our conclusion that active regions may not be heated by Alfv\\'en waves. These results seem to point to an important role of Alfv\\'en wave heating in Quiet Sun regions, where loops are often long, expand more than in active regions, and kG (or higher) bright points are ubiquitous. Another important result we have found is that both long (above 100 s) and short period (below 50 s) waves (where the period is the resulting period of the longitudinal modes from mode conversion; the period of the Alfv\\'en waves is twice the stated period) play an important role in the heating of the loops. Long period waves produce very hot heating events in the corona and thus increase the average temperature of the corona due to the strong shocks they produce. Short period waves are responsible for keeping the corona, thus making it stable through uniform heating from the numerous weak shocks they produce. We have therefore a compromise between long and short period waves leading to efficient Alfv\\'en wave heating. The 300 s period waves are found to have considerable power to heat the corona even with low photospheric velocity amplitudes of $\\sim0.2$ km s$^{-1}$. For these waves resonances seem to be triggered in the loop, which allow higher transmission from the waves into the corona and subsequent mode trapping leading to efficient dissipation. This resonant damping mechanism is however not observed for random wave generation (white noise spectrum) or with other geometric parameters for the loop. A thorough study of these coronal resonances for Alfv\\'en waves will be the subject of a future paper." }, "0910/0910.2226_arXiv.txt": { "abstract": "{Recent progress in instrumentation enables solar observations with high resolution simultaneously in the spatial, temporal, and spectral domains. } % {We use such high-resolution observations to study small-scale structures and dynamics in the chromosphere of the quiet Sun. } % {We analyse time series of spectral scans through the \\ion{Ca}{II}~854.2~nm spectral line obtained with the CRISP instrument at the Swedish 1-m Solar Telescope. The targets are quiet Sun regions inside coronal holes close to disc-centre. } % {The line core maps exhibit relatively few fibrils compared to what is normally observed in quiet Sun regions outside coronal holes. The time series show a chaotic and dynamic scene that includes spatially confined ``swirl'' events. These events feature dark and bright rotating patches, which can consist of arcs, spiral arms, rings or ring fragments. The width of the fragments typically appears to be of the order of only 0\\,\\farcs2, which is close to the effective spatial resolution. They exhibit Doppler shifts of $-2$ to $-4$\\,km$/$s but sometimes up to $-7$\\,km$/$s, indicating fast upflows. The diameter of a swirl is usually of the order of 2\\,\\arcsec. At the location of these swirls, the line wing and wide-band maps show close groups of photospheric bright points that move with respect to each other. } % {A likely explanation is that the relative motion of the bright points twists the associated magnetic field in the chromosphere above. Plasma or propagating waves may then spiral upwards guided by the magnetic flux structure, thereby producing the observed intensity signature of Doppler-shifted ring fragments. } ", "introduction": "Advances in observational performance have revolutionised our understanding of the solar chromosphere over the last years. It is in particular the combination of high spectral, temporal, and spatial resolution that let us discover details and phenomena hitherto unaccessible. Our picture of the solar atmosphere changed from a static plane-parallel stratification to a very complex compound of intricately coupled dynamic domains (see recent reviews by e.g. Schrijver 2001; Judge 2006; Rutten 2006, 2007;Wedemeyer-B{\\\"o}hm et al. 2009). And yet many details concerning the chromosphere and its connection to the layers above and below remain elusive. The progress in our understanding is certainly hampered by the complexity and accessibility of currently available diagnostics probing that atmospheric layer. Among them are the spectral lines of \\ion{Ca}{II}, which are formed over a very extended height range in the atmosphere. Only the line cores truly originate from above the photosphere. Observing the line core exclusively therefore requires filters with a very narrow transmitted wavelength range, as the width of the line cores is only of the order of 100\\,pm or less. Considering the spectral, temporal, and spatial domains, instrumental limitations enforce compromises which until recently only allowed high resolution in one or two of the domains, at the cost of the remaining one(s). This situation has now improved substantially with the installation of fast two-dimensional spectropolarimetric imagers at solar telescopes with large aperture. Examples are the IBIS instrument (Cavallini 2006) at the Dunn Solar Telescope, CRISP (Scharmer et al. 2008) at the Swedish Solar Telescope (SST), and the G{\\\"o}ttingen Fabry-P{\\'e}rot (Puschmann et al. 2006) Achieving high resolution simultaneously in all these three domains is like opening a new observing window on the chromosphere. Here we report on the discovery of small but fast rotating swirls in the chromosphere. There are quite a few examples known of rotating or vortex-like motions on the Sun: on large scales in the form of rotating sunspots (e.g., Brown et al. 2003), on smaller scales like vortices in the photospheric granulation (Brandt et al. 1988), and on the smallest scales in the form of whirlpool motion by magnetic bright points in the inter-granular lanes (Bonet et al. 2008). This rotating motion receives considerable interest since these photospheric motions have a profound effect on the outer-atmospheric magnetic fields as they have their roots in the photoshere. The stresses that are built up through rotation of magnetic fields are linked to the onset of solar flares and coronal heating (e.g., Parker 1983). After this introduction we describe the observations in Sect.~\\ref{sec:observ}, which are analysed in Sect.~\\ref{sec:ana}. A discussion and conclusions are given in Sect.~\\ref{sec:discus}. ", "conclusions": "\\label{sec:discus} At first glance, the swirls in the coronal hole are reminiscent of the convectively driven vortex flows reported by Bonet et al. (2008). In a high-resolution SST G-band filtergram time sequence, they found examples of groups of BPs displaying clear ``whirlpool'' motion. Bonet et al. associate the photospheric whirlpools with vigorous downdrafts at the vertices of intergranular lanes that have been predicted by convection models. The G-band BPs act as flow tracers which follow spiral trajectories when they are engulfed by the downdrafts. The size of these whirlpools is less than an arcsecond. Contrastingly, the photospheric BPs associated with the chromospheric swirls in our data do not follow clear spiral trajectories. It is therefore unclear if photospheric whirlpools and chromospheric swirls are connected. Our wide-band images usually show a small number of photospheric BPs that move in relation to each other, buffeted by the lateral granular flows at the top of the convection zone. Being confined to the intergranular lanes, the BPs can eventually come close and even move past each other very closely. Our interpretation is that these close encounters result in a deformation and twist of the magnetic field continuing above the BPs. While tending to unwind the build-up stresses, gas may spiral upwards, producing the observed intensity signature of Doppler-shifted ring fragments. Such a process may be impeded by the presence of a stronger magnetic field component in the chromosphere in the form of a ``magnetic canopy'' (e.g., Schrijver \\& Title 2003) as it exists above quiet Sun regions outside coronal holes. Swirls are therefore expected to be significantly less frequent in quiet Sun regions outside coronal holes, which have a high fibril coverage. Quantifying this assumption is unfortunately difficult as the detection of swirls in Ca line core maps is complicated by the presence of fibrils. Events ongoing in the layer just below the fibrils could be obscured by the horizontal component of the magnetic field (Vecchio et al. 2009). The exact cause for the swirl phenomenon remains to be found. The magnetic fields, which are supposedly responsible for forcing the rotating plasma or propagating waves on ring-like or spiral trajectories, are not directly detected in our CRISP observations. However, all our detected swirls show a small group of BPs in the centre, implying a causal connection. We interpret these chromospheric swirl motions and associated BP motions as a direct indication of upper-atmospheric magnetic field twisting and braiding as a result of convective buffeting of magnetic footpoints. This mechanism is one of the prime candidates for coronal heating (Parker 1988)." }, "0910/0910.0015_arXiv.txt": { "abstract": "In a recent paper we presented the first semi-analytic model of galaxy formation in which the Thermally-Pulsing Asymptotic Giant Branch phase of stellar evolution has been fully implemented. Here we address the comparison with observations, and show how the TP-AGB recipe affects the performance of the model in reproducing the colours and near-IR luminosities of high-redshift galaxies. We find that the semi-analytic model with the TP-AGB better matches the colour-magnitude and colour-colour relations at $z \\sim 2$, both for nearly-passive and for star-forming galaxies. The model with TP-AGB produces star-forming galaxies with red V-K colours, thus revising the unique interpretation of high-redshift red objects as 'red \\& dead'. We also show that without the TP-AGB the semi-analytic model fails at reproducing the observed colours, a situation that cannot be corrected by dust reddening. We also explore the effect of nebular emission on the predicted colour-magnitude relation of star-forming galaxies, to conclude that it does not play a significant role in reddening their colours, at least in the range of star-formation rates covered by the model. Finally, the rest-frame K-band luminosity function at $z \\sim 2.5$ is more luminous by almost 1 magnitude. This indicates that the AGN feedback recipe that is adopted to regulate the high-mass end of the luminosity function should be sophisticated to take the effect of the stellar populations into account at high redshifts. ", "introduction": "Hierarchical clustering is the favoured scenario to describe the formation and evolution of matter structures in the universe (White \\& Rees, 1978), and semi-analytic models of galaxy formation proved themselves to be a powerful tool of investigation since the first formulation (White \\& Frenk, 1991). Over the years, many such models have been developed (see for instance Balland et al. 2003, Baugh et al. 2005, Bower et al. 2006, Cattaneo et al. 2008, Cole et al. 2000, De Lucia et al. 2004, Hatton et al. 2003, Kauffmann et al. 1993, Menci et al. 2006, Monaco et al. 2007, Somerville et al. 2008). The successes and failures of these models are strictly linked to those of the hierarchical scenario itself, ultimately depending on the mechanisms of mass accretion of objects as a function of time. The large-scale structure and the integrated properties of the galaxy population (such as the total stellar mass density for instance) are well reproduced. The detailed evolution of galaxies however presents several puzzling aspects, such as e.g. the size of the disks of spirals (Burkert 2009), or the $\\alpha$-enhancement and the ages of the stellar populations in massive ellipticals (Thomas 1999, Thomas et al. 1999, Cimatti et al. 2004, Nagashima et al. 2005, Thomas et al. 2005, Pipino et al. 2009, Kormendy et al. 2009). Moreover, global properties of the galaxy population such as the evolution of the stellar mass function (Cole et al. 2001, Bell et al. 2003, Bundy et al 2006-2009, Pozzetti et al. 2009, Colless et al. 2007) are still not reproduced in the models (Bundy et al. 2007, Marchesini et al. 2009, Kajisawa et al. 2009, Kodama \\& Bower 2003), although there is controversy on this point (Drory et al. 2004, Benson et al. 2007). It is debated if any of these problems can possibly be resolved with enhanced resolution in the simulations and more sophisticated recipes in the models. One of the most problematic issues for the models is to reproduce the abundance of high redshift luminous galaxies (e.g., Conselice et al. 2007, Cimatti et al 2004, van Dokkum et al. 2004, 2006). This difficulty is partly due to a mis-interpretation of the nature of these objects. The high-luminosity end of the galaxy population up to redshift $z \\sim 2.5$ consists in fact of objects that look like the early-type galaxies in the local universe, \\textit{i.e.} they are characterized by very red colours in the optical and near-IR, and high near-IR luminosities (Mancini et al. 2009, Cimatti et al 2004, McCarthy et al. 2004, Daddi et al. 2005, Saracco et al. 2005, Kriek et al. 2006). Local ellipticals showing the same photometry are old (with stellar populations older than $\\sim1$ Gyr), passively evolving, and with stellar masses $M_{star} > 10^{11} \\ M_{\\odot}$. Moreover, newer observations are building the case for the presence of extremely red and IR-luminous objects at even higher redshifts (Rodighiero et al. 2007, Mancini et al. 2009, Fontana et al. 2009). The problem posed by the presence of these high redshift ($z>2$) red and luminous galaxies stems from the consensus that they are massive objects evolving passively, the so-called 'red \\& dead' galaxies. With the stellar population models currently used in the semi-analytic models in the literature (for the most part Bruzual \\& Charlot 2003, hereafter BC03), the only way to explain the high near-IR luminosities of high-redshift galaxies is to advocate very high galaxy masses and very old ages of the stellar populations. But these are not achieved in the actual model realizations (except at low redshifts), because the hierarchical mass assembly has an intrinsic difficulty in putting together massive and old objects at early epochs. In fact, the hierarchical scenario predicts a steady decline of the abundance of massive galaxies with increasing redshift (van Dokkum et al. 2004). In Tonini et al. (2009) we showed that the predictions of colours and luminosities of galaxies at high redshift in a semi-analytic model are greatly affected by the recipes in use for the stellar populations, expecially the inclusion of the Thermally-Pulsing Asymptotic Giant Branch (TP-AGB). As shown in M05, in stellar populations of intermediate age ($\\leq 0.2 - 2$ Gyr) the TP-AGB phase dominates the near-IR luminosity, with a contribution up to $80 \\%$ in the rest-frame K band, and contributes to up to $40 \\%$ of the bolometric luminosity (M05). High-redshift galaxies, in which the mean age of the stellar populations is in that range, are expected to be dominated by the TP-AGB emission in the near-IR. This has been recently confirmed by SED-fitting of observations made with the Spitzer Space Telescope (Maraston et al. 2006, Cimatti et al. 2008). In Tonini et al. (2009) we included a complete treatment of the TP-AGB phase in the semi-analytic model GalICS (Hatton et al. 2003), by implementing the M05 stellar population models into the code, and showed that the rest-frame $V-K$ colours at high redshift get redder by more than 1 magnitude. Relatedly, the K-band mass-to-light ratio is shifted towards luminosities 1 magnitude higher for a given galaxy mass. Notably, actively evolving, star-forming high-redshift galaxies are predicted to have $V-K$ colours and near-IR luminosities similar to those of local, passively evolving massive systems (Tonini et al. 2009). Once the stellar emission is correctly modeled with an exhaustive treatment of all the significant phases of stellar evolution, a more accurate comparison between the semi-analytic model predictions and the data is possible. In particular, the performance of the semi-analytic model in reproducing the observed colours and luminosities in the near-IR becomes meaningful to test the hierarchical mass assembly at different redshifts. In the literature the comparison between galaxy formation models and data is typically done by obtaining physical properties for the real objects through application of stellar population models to data. However, this approach carries several degeneracies, including the adopted population synthesis model, the recipe for star formation history, the choice of metallicity, etc. When a realistic errorbar including all these variables ia attached to the observationally derived quantity, such as in Marchesini et al. 2009 (and see also Conroy et al. 2009 for a discussion), the results of such comparisons may not be clear-cut. In this paper we adopt a different philosophy for the comparison between model and data. Instead of using processed data in the rest-frame system, we consider raw, unprocessed, apparent magnitudes straight out of the catalogues. We then produce mock catalogues out of the simulation, so that the output spectra of the model galaxies are redshifted in the observer's frame. The model apparent magnitudes and colours can then be directly compared with the observational data. This comparison yields direct information about the physical quantities in the model in use. This procedure is straightforward and does not add substantial degeneracy that can jeopardize the comparison. A degeneracy that clearly remain is how dust reddening affects the intrinsic stellar emission, as recently pointed out by Guo \\& White (2008) and Conroy et al. (2009). However, we shall show that considering the intrinsic star formation rates in the model and using data mapping the rest-frame near-IR, such an uncertainty plays actually a minor role. The structure of the paper is as follows. In Section 2 we briefly introduce the new semi-analytic model GalICS with the TP-AGB implementation through the M05 models (as from Tonini et al. 2009). In Section 3 we describe the data samples used for our analysis. In Section 4 we compare the colour-magnitude and colour-colour relations predicted by the model against samples of $z \\sim 2$ galaxies. In Section 4 we compare the model rest-frame K-band luminosity function in the M05 and Pegase cases with the predictions by other semi-analytic models. In Section 5 we discuss our results. ", "conclusions": "In a recent work (Tonini et al. 2009) we introduced the complete treatment of the TP-AGB phase of stellar evolution into a semi-analytic model of galaxy formation, by inserting the Maraston (2005) SSP models into the code GalICS. In the work presented here we compared the predictions on the near-IR luminosities and colours of high-redshift galaxies with data samples of nearly-passive and star-forming galaxies around $z \\sim 2$. Our main results are: $\\cdot$ the TP-AGB is fundamental to allow the semi-analytic model to reproduce the observed optical and near-IR colours of both nearly-passive and star-forming galaxies at $z \\sim 2$; the inclusion of the TP-AGB increases the Irac3 luminosity (rest-frame K) and shifts the H-Irac3 (rest-frame V-K) colours by more than 1 magnitude; $\\cdot$ without the TP-AGB, it is not possible to match the observed galaxy colours and luminosities by a modification of the dust reddening recipe alone; $\\cdot$ the TP-AGB emission does not alter the optical luminosity and colours of star-forming galaxies. On the other hand, star formation does not dilute the TP-AGB emission in the near-IR. Even star-forming galaxies, very blue in the optical, can be very red in the near-IR. Therefore the labelling of red galaxies as 'red and dead' is misleading; $\\cdot$ the nebular emission, produced by young stellar populations, does not add a significant contribution to the colours of star-forming galaxies, in the range of star-formation rates covered by the model; for SFR$>70 \\ M_{\\odot}/yr$ the rest-frame V-K colour is reddened by 0.2-0.4 mags; $\\cdot$ the predicted mass-luminosity relation is affected by the inclusion of the TP-AGB; for a given galaxy mass, the rest-frame K-band luminosity is higher by more than 1 mag at $z>1$. As a consequence, the K-band luminosity function predicted by the model with the TP-AGB shifts redwards, expecially at the high-mass end, for $z>1$ (by $\\sim 0.7$ mag at $z \\sim 2.5$). The spread in the luminosity function between runs with and without the TP-AGB is comparable to the scatter caused by different AGN-feedback recipes in the literature. Note that the high-mass end of the luminosity function in the near-IR is dominated by spheroids, or the progenitors of today's spheroids. If the use of the TP-AGB in the semi-analytic model shifts the luminosity function by $\\sim$1 mag at the high-mass end, it means that the mass-to-light ratio is lower by a factor of $\\sim$2.5 for a given luminosity. When galaxy masses are inferred from observations by the use of these models, they are lower by the same factor (as shown in M05). This may rise the question of whether there is enough mass in spheroids at high redshift to account for the $\\sim 50\\%$ of stellar mass in ellipticals measured in the local universe. However, the model correctly predicts the stellar mass density at all redshifts, meaning that only the \\textit{distribution of galaxy masses} is at tension with observations, \\textit{if the TP-AGB is not taken into account}. In fact, hierarchical models in general predict a faster evolution of the high-mass end of the stellar mass function than currently inferred from observations (see for instance Conselice et al. 2007). A more accurare derivation of galaxy masses through complete stellar population models with the TP-AGB, coupled with more accurate predictions from hierarchical models with the right input SSP, surely contribute to alleviate the discrepancy. The inclusion of the TP-AGB allows the semi-analytic model to reproduce the very red end of the galaxy population at $z \\sim 2$, both for nearly-passive and for star-forming objects. It allows the model to do so with a comfortable range of galaxy masses and dust reddening. Most importantly, it contributes to a realistic and comprehensive treatment of the galaxy light emission in galaxy formation models, making them a much more precise tool to test our understanding of galaxy assembly. The implementation of the TP-AGB allows the model to produce, at a given stellar mass, redder and more luminous galaxies in the near-IR, expecially at high redshift where the ages of the stellar populations peak around the epoch of maximal emission from this stellar phase. In case of nearly-passively evolving galaxies, the model can reproduce the red colours and high K-band magnitudes without invoking too large stellar masses or too old ages, which would be problematic in the hierarchical context. In the case of star-forming galaxies, the TP-AGB still increases the near-IR luminosity and makes the galaxies redder, without offsetting the blue optical colours. Thus, observed red colours in the near-IR do not necessarily imply old ages and passive evolution, a fact that again would be problematic for the hierarchical picture at high redshift. In general, the introduction of the TP-AGB in the models is a step forward in reconciling the hierarchical assembly mechanism with the observations of the high-redshift universe." }, "0910/0910.2821_arXiv.txt": { "abstract": "We present a sensitive 870\\,\\micron\\ survey of the Extended Chandra Deep Field South (ECDFS) combining 310 hours of observing time with the Large Apex BOlometer Camera (LABOCA) on the APEX telescope. The LABOCA ECDFS Submillimetre Survey (LESS) covers the full $30'\\times30'$ field size of the ECDFS and has a uniform noise level of $\\sigma_{870\\mu{\\rm m}}\\approx1.2$\\,mJy\\,beam$^{-1}\\,$. LESS is thus the largest contiguous deep submillimetre survey undertaken to date. The noise properties of our map show clear evidence that we are beginning to be affected by confusion noise. We present a catalog of 126 submillimetre galaxies (SMGs) detected with a significance level above 3.7\\,$\\sigma$, at which level we expect 5 false detections given our map area of 1260 arcmin$^2$.\\\\ The ECDFS exhibits a deficit of bright SMGs relative to previously studied blank fields but not of normal star-forming galaxies that dominate the extragalactic background light (EBL). This is in line with the underdensities observed for optically defined high redshift source populations in the ECDFS (BzKs, DRGs, optically bright AGN and massive K-band selected galaxies). The differential source counts in the full field are well described by a power law with a slope of $\\alpha=-3.2$, comparable to the results from other fields. We show that the shape of the source counts is not uniform across the field. Instead, it steepens in regions with low SMG density. Towards the highest overdensities we measure a source-count shape consistent with previous surveys. The integrated 870\\micron\\ flux densities of our source-count models down to $S_{870\\mu{\\rm m}}=0.5$\\,mJy account for $>65\\%$ of the estimated EBL from {\\it COBE} measurements. We have investigated the clustering of SMGs in the ECDFS by means of a two-point correlation function and find evidence for strong clustering on angular scales $<1'$ with a significance of $3.4\\,\\sigma$. Assuming a power law dependence for the correlation function and a typical redshift distribution for the SMGs we derive a characteristic angular clustering scale of $\\theta_0=14''\\pm7''$ and a spatial correlation length of $r_0=13\\pm6\\,h^{-1}$\\,Mpc. ", "introduction": "One of the most significant findings of the IRAS survey was the identification of a population of ultraluminous infrared galaxies (ULIRGs) that emit the bulk of their bolometric luminosity at far-IR wavelengths \\citep{sanders96}. Surveys in the submillimetre and millimetre wavebands over the past decade have shown that ULIRGs are much more common at high redshift compared to the local universe \\citep[e.g.][]{barger99,cowie02,borys03,webb03,greve04,laurent05, coppin06,pope06,bertoldi07,beelen08,knudsen08,scott08,austermann09}. These surveys show that the comoving volume density of luminous submillimetre galaxies (SMGs) increases by a factor of $1000$ out to $z\\sim 2$ \\citep{chapman05}. Therefore luminous obscured galaxies at high redshift could dominate the total bolometric emission of galaxies at those epochs \\citep{blain99,lefloch05} The identification and study of submillimetre galaxies has proved challenging since their first detection. The limited mapping speed of typical (sub)millimetre bolometer cameras meant that only few very bright examples have been found, although gravitational lensing initially aided somewhat \\citep[e.g.][]{smail97,ivison98}. Attempts to map large fields at submillimetre wavelengths have involved the use of patchworks of small ''jiggle'' maps \\citep[e.g.][]{coppin06} or mixtures of single-bolometer photometry, small jiggle maps and shallow scan maps used to construct a ''Super-map'' of GOODS-N \\citep{borys03,pope06}. Both of these approaches raise concerns about the homogeneity of the resulting maps and hence the reliability of the resulting source catalogues. Scan maps, where the array is continuously moved on the sky to trace out a closed pattern, should result in much more homogeneous coverage and mapping, while at the same time allowing for a reliable removal of the bright emission from the atmosphere. This technique has been used at submillimetre and millimetre wavelengths \\citep[e.g. at 350$\\mu$m,][at 1100$\\mu$m]{kovacs06,austermann09}, however, no deep survey have employed such a technique in the 870-$\\mu$m window where most of the published work on SMGs has been undertaken. Drawing this distinction between 870-$\\mu$m and 1100-$\\mu$m surveys may appear surprising given the modest difference between the two wavelengths and the assumed unstructured nature of the dust spectrum at these wavelengths. Despite only a 25\\% difference in the two wavelengths, there are hints of significant differences in the populations identified at 870-$\\mu$m and 1100-$\\mu$m \\citep[e.g.][]{greve04,younger08}, although these may in turn reflect the different mapping techniques used in individual studies. The advent of the new Large APEX Bolometer Camera \\citep[LABOCA,][]{siringo09}, with an instantaneous 11.4$'$ field of view, on the 12-m APEX telescope \\citep{guesten06} provided the opportunity to undertake the first sensitive and uniform panoramic survey of the extragalactic sky at 870$\\mu$m. To exploit this opportunity a number of groups within the Max Planck Gesellschaft (MPG) and the European Southern Observatory (ESO) communities proposed a joint public legacy survey of the Extended {\\it Chandra} Deep Field South (hereafter ECDFS) to the MPG and ESO time allocation committees: the LABOCA ECDFS submillimetre survey (LESS). The ECDFS covers a 0.5$^\\circ\\times$\\,0.5$^\\circ$ region centered on the {\\it Chandra} Deep Field South (CDFS) at RA $03^h32^m28^s.0$ Dec $-27^{\\circ}48^{'}30^{''}.0$. This field has very low far-infrared backgrounds and good ALMA visibility and hence has become one of the pre-eminent fields for cosmological survey science. As a result, the ECDFS is unique in the Southern Hemisphere in the combination of area, depth and spatial resolution of its multiwavelength coverage from X-rays through optical, near- and mid-infrared to the far-infrared and radio regimes. The central part of this field is coincident with the CDFS \\citep{giacconi02} which has now reached a depth of 2\\,Ms \\citep{luo08} and the deep {\\it Hubble Space Telescope} ({\\it HST}) imaging of the GOODS-S field \\citep{giovalisco04} and the {\\it Hubble} Ultra Deep Field \\citep[UDF,][]{beckwith06}. In addition to the extremely deep observations of the central regions of this field as part of the CDFS, GOODS and {\\it Hubble} Ultra Deep Field surveys, the full 0.5$^\\circ$ field has extensive multiwavelength imaging available including: 250-ks {\\it Chandra} integrations over the whole field \\citep{lehmer05}; deep and multi-band optical imaging by COMBO-17 \\citep[][2008]{wolf04} and MUSYC \\citep{gawiser06} including {\\it HST} imaging for the GEMS project \\citep{caldwell08}; near-infrared imaging by MUSYC \\citep{taylor09}; deep mid-infrared imaging with IRAC as part of SIMPLE \\citep{damen09} and using the MIPS instrument at 24, 70 and $160\\mu$m by FIDEL (Dickinson \\etal\\ in prep.). Longer wavelength coverage comes from BLAST \\citep{devlin09} at 250, 350 and $500\\mu$m (and in the near future from {\\it Herschel}), while radio coverage of this field is reported by \\citet[][]{miller08} and \\citet[][]{ivison09}. The LABOCA survey of the ECDFS adds a waveband that pin-points the thermal emission from luminous dusty galaxies at $z\\sim1$--8: a powerful addition to this singularly well-studied region -- ideally placed for VLT observations and early science follow-up with ALMA. The completed LESS project provides a representative, homogeneous and statistically-reliable sample of the SMGs with the high-quality, multiwavelength data required to yield identifications, constrain their redshifts, bolometric luminosities and power sources and hence determine their contribution to the total star formation density at high redshift. These sources can be related in unprecedented precision to other populations of AGN and galaxies within the same volume to understand the place of SMGs in the formation and evolution of massive galaxies at high redshift. The survey is also sufficiently large that it should also yield examples of rare classes of SMGs, such as very high redshift sources, $z>4$ \\citep{coppin09}. These same data also provide submillimetre coverage of large numbers of high-redshift galaxies and AGN to determine their bulk submillimetre properties from the stacking analysis of sub-samples as a function of population, redshift, environment, etc. \\citep{greve09,lutz09}. Together, these two techniques allow us to sample two orders of magnitude in bolometric luminosity -- from hyperluminous infrared galaxies with $10^{13}$\\,L$_\\odot$ which are directly detected in the LABOCA maps, down to luminous infrared galaxies at $10^{11}$\\,L$_\\odot$ which are detected statistically through stacking. This range in luminosity encompasses the variety of populations expected to dominate the bolometric emission at $z \\sim 1$--3 and the cosmic submillimetre background. In this paper we present a detailed description of the observations, reduction and analysis of the LABOCA observations of the ECDFS and the resulting catalogue of submillimetre galaxies. The observations are described in \\S2, \\S3 presents our results and we discuss these in \\S4. Finally, in \\S5, we give our summary and the main conclusions of this work. We assume a cosmology with $H_0=70$\\,km\\,s$^{-1}$\\,Mpc$^{-1}$, $\\Omega_\\Lambda=0.7$ and $\\Omega_M=0.3$. ", "conclusions": "We have presented a deep 870\\,\\micron\\ survey of the ECDFS using LABOCA on the APEX telescope at Llano de Chajnantor in Chile. This is the largest contiguous deep submm survey to date. Our map has a highly uniform noise level across the full $30'\\times30'$ field of 1.2\\,mJy$\\,{\\rm beam}^{-1}$ and our survey is $>95\\%$ complete for sources down to a flux limit of 6.5\\,mJy. Our main findings are summarized as follows: \\begin{itemize} \\item At the (beam smoothed) spatial resolution of $27''$ of our survey we find that the map's noise level is affected by confusion noise arising from faint, individually undetected SMGs. From the rms noise as a function of integration time we derive a confusion noise of $\\sigma_{\\rm c}\\approx\\, 0.9\\,{\\rm mJy}\\,{\\rm beam}^{-1}$. \\item We identify 126 submm sources in a search area of 1260\\,arcmin$^2$ above a signal to noise threshold of $3.7\\,\\sigma$, which corresponds to an expected false detection rate of 5 sources. \\item We have determined the differential and integrated source counts using a $P(D)$ analysis and an estimate based on our source catalog. Both results are in reasonable agreement and show that SMGs in the ECDFS are underabundant by a factor of $\\sim2$ for sources brighter than 3\\,mJy compared to the average of previous surveys. Under the assumption that the bulk of the sources are at $z>0.5$, this implies an underdensity of ULIRGs with $\\lfir>2\\times10^{12}\\lsol$ compared to other blank fields that have been observed in the submm. The source counts are well described by a single power law with a slope of $\\alpha=3.2\\pm0.2$. \\item We derive the angular two-point correlation function for the SMGs and find clustering on angular scales $<1'$ with a significance up to $3.4\\,\\sigma$. Assuming a power law dependence for the correlation function we derive a clustering amplitude of $A_w=0.011\\pm0.0046$ or a characteristic angular scale of $\\theta_0=14''\\pm7''$ for $\\gamma =1.8$. Assuming a redshift distribution similar to that observed for spectroscopically confirmed SMGs, we derive a correlation length of $r_0=13\\pm6\\,h^{-1}$\\,Mpc, somewhat larger than previous estimates of the 3-D clustering of SMGs but in agreement with the clustering derived for 24\\,\\micron\\ selected ULIRGs. \\item We have investigated for the first time the spatial variations of the SMG source counts. We find that the differential source counts in regions with an overdensity of SMGs have a different shape compared to those with underdensities. While the counts in underdense regions are well fitted by a single power law with a slope of $\\alpha=3.6\\pm0.3$, the counts in the overdensities are significantly shallower with $\\alpha=2.9\\pm0.2$. The counts in the overdensities are slightly better described by a broken power law or a Schechter function. For flux densities below 8\\,mJy we find a slope of $\\alpha=2.4\\pm0.15$, for sources above this limit the counts are much steeper with $\\beta=4.7\\pm0.6$. This may indicate an intrinsic turn-over in the underlying luminosity function placing an upper limit on the FIR luminosity. \\item The integrated 870\\,\\micron\\ flux density derived from our survey is $>29-32$\\,Jy\\,deg$^{-2}$ for sources brighter than $\\sim0.5$\\,mJy which corresponds to $>65-70$\\% of the extragalactic background light estimated from {\\it COBE} measurements. We do not find a significant difference of the quantity between SMG over- and underdensities. We conclude that ECDFS is underabundant of ULIRGs but not of more typical star forming systems with lower FIR luminosities, which dominate the extragalactic background light. \\end{itemize}" }, "0910/0910.1539_arXiv.txt": { "abstract": "We show that simple models of scalar-field dark energy leave a generic enhancement in the weak-lensing power spectrum when compared to the $\\Lambda$CDM prediction. In particular, we calculate the linear-scale enhancement in the convergence (or cosmic-shear) power spectrum for two best-fit models of scalar-field dark energy, namely, the Ratra-Peebles and SUGRA-type quintessence. Our calculations are based on linear perturbation theory, using gauge-invariant variables with carefully defined adiabatic initial conditions. We find that geometric effects enhance the lensing power spectrum on a broad range of scales, whilst the clustering of dark energy gives rise to additional power on large scales. The dark-energy power spectrum for these models are also explicitly obtained. On degree scales, the total enhancement may be as large as $30$-$40\\%$ for sources at redshift $\\sim1$. We argue that there are realistic prospects for detecting such an enhancement using the next generation of large telescopes. ", "introduction": "Over the past years, a concordance picture of our universe has emerged from a number of cosmological probes. In this paradigm, the dominant form of matter is cold and dark (CDM), but most of the mass-energy budget is in the form of dark energy, manifest in the accelerated expansion of the universe [for some of the earliest evidence from Type Ia supernovae, see for example \\cite{GoldPerl1998} and \\cite{Riess1998Aj}]. Some of the key goals in cosmology from the analysis of upcoming high-precision experiments within the next couple of decades are to determine the nature of dark energy, and whether there is a tension with its interpretation as a cosmological constant. A decisive way to distinguish whether dark energy is the cosmological constant or some other dynamical entity is to establish whether the dark-energy equation-of-state parameter $w\\sub{DE}$ is constant or varies with redshift. The variation of $w\\sub{DE}$ can be probed via a range of astronomical techniques, including measurements of light-curves of type Ia supernova (SNIa), cosmic microwave background (CMB) anisotropies, galaxy redshift surveys, galaxy cluster abundance and cosmological weak lensing (see \\cite{copeland,frieman+} for a summary of the various techniques). Out of these techniques, weak lensing holds great promise in the understanding of dark energy, and has been identified as the most powerful individual technique, or most important component in a multi-technique study, provided systematic errors are well understood \\citep{detf, peacock}. Its utility as a dark-energy probe comes from sensitivity to both the expansion history of the universe and to the rate of growth of structures. Observation of weak lensing is still in its infancy, with the first detection of cosmic shear made only nine years ago \\citep{wittman}. At present, there are a number of upcoming ambitious experiments - both terrestrial [\\eg LSST, ELT \\citep{lsst, elt}] and space-based [\\eg JDEM\\footnote{http://jdem.gsfc.nasa.gov}, Euclid\\footnote{http://sci.esa.int/euclid}], aimed at measuring cosmic shear to high accuracy. Given this rapid development, it is worth investigating if the upcoming weak-lensing experiments could discriminate simple models of dynamical dark energy from the cosmological constant. Many previous works on this problem concentrate solely on the geometrical effects of dynamical dark energy. Furthermore, many authors assume dynamical dark energy with $w\\sub{DE}=$ constant, or some algebraic expression [numerous examples may be found the review by \\cite{copeland}]. Whilst this is a useful approximation for studying cosmological dynamics at late-time, it is impossible to extrapolate these simple forms of $w\\sub{DE}$ to early times, unless some \\ii{ad hoc} mechanisms are invoked. As a result, one cannot consistently analyse the perturbations in dark energy, and many authors then neglect these perturbations altogether. Even those who do include dark-energy perturbations often gloss over the issue of initial conditions, and simply assume that the perturbations are insensitive to initial conditions without any justification. In this work, we shall calculate the effects of dynamical dark energy on weak-lensing measurements without making any of the assumptions in the previous paragraph. We shall analyse simple models of dynamical dark energy using gauge-invariant perturbation theory, with well-defined initial conditions. As we shall see, these models leave some interesting imprints on the weak-lensing observables on all scales. Ultimately, we shall give a basic assessment of the prospects for distinguishing dynamical dark energy from the cosmological constant using weak-lensing measurements. One particularly important aspect of this work is the inclusion of dark-energy perturbations throughout our analysis. There have been a number of recent works examining the effects of dark-energy perturbations on cosmological observables \\citep{dent, sapone,hwang2}, including some interesting numerical simulations \\citep{alimi, jennings}. Our approach is complementary to these works, and goes further in quantitatively establishing the effects of dynamical dark energy on weak-lensing power spectra. As a by-product, we obtain explicitly, for the first time, the dark-energy power spectrum for these models. Such a spectrum is useful as it quantifies the clustering of dark energy as a function of physical length scale. Throughout this paper, we work in Planck units in which $c=\\hbar=1$. ", "conclusions": "In this work, we have investigated whether there are observable signatures of dynamical dark energy in weak-lensing measurements. Our basic assumptions are that the background cosmology is a flat FRW universe, containing radiation, dark matter and the simplest type of scalar-field dark energy (with no hot dark matter, baryons or tensor modes). Our technique involves calculating the linear density power spectra by integrating six coupled differential equations, expressed in terms of gauge-invariant density and velocity variables for dark matter, radiation and dark energy (equations \\ref{d1}-\\ref{v3}). We assume adiabatic initial conditions for all three types of perturbations, and normalise the matter power spectrum (with primordial form $\\sim k^{n_s}$) on CMB scales. We base our analysis specifically on two simple models of quintessence, namely the Ratra-Peebles and SUGRA-type potentials, with model parameters that best fit the recent CMB and combined supernova data. We find that the best-fit quintessence models yield matter power spectra that are hardly distinguishable from that of the $\\Lambda$CDM model (figure \\ref{figpm}) except on large scales where dark-energy clustering is strongest, as shown in figure \\ref{figde}. We believe this is the first time that the dark-energy power spectra for these oft-cited models are explicitly calculated. When the weak-lensing signals are extracted from the 3D spectra, we found an imprint in the form of an all-scale enhancement in the convergence (or the cosmic shear) power spectrum compared with $\\Lambda$CDM cosmology. For sources at redshift 0.5, for instance, the enhancement may be as large as $40\\%$ at about $2^\\circ$ scale (or 20\\% at $z=1$). These enhancements can be attributed to two main reasons, namely i) the change in distance-redshift relation (figure \\ref{figdistance}), and ii) the lensing by large-scale dark-energy clusters (figure \\ref{figignore}). Assuming that the primary sources of errors in the weak-lensing measurements will be from cosmic variance and intrinsic ellipticity dispersion in the shapes of distant galaxies, there are good prospects for detecting this enhancement for sources at redshift $z\\sim0.5$ using future facilities with surveys covering most of the sky, and capable of measuring ellipticities for hundreds of galaxies per arcmin$^{2}$ (figure \\ref{figerrors}). On the timescale of a decade, one class of telescope capable of such a survey is the E-ELT, 42m in diameter and including adaptive optics, for which operations are planned to start in 2018. To put this requirement in context, current ground-based surveys typically measure ellipticities for a few tens of galaxies per arcmin$^{2}$ over order 100 square degrees (e.g. CFHTLS-Wide\\footnote{http://terapix.iap.fr}), and the highest specification near-future survey DES\\footnote{http://www.darkenergysurvey.org} will measure a comparable number density of galaxies over 5000 square degrees. Here we restricted our consideration to distant galaxies at the same redshift, but for a real survey with sources distributed in redshift, photometric redshift estimates can be obtained using observations in multiple filters. Tomographic information using the auto- and cross-correlations between the cosmic shear in various redshift bins, will give additional leverage in distinguishing between models of dark energy. \\bbb \\no{\\bf" }, "0910/0910.3827_arXiv.txt": { "abstract": "{Counting clusters is one of the methods to constrain cosmological parameters, but has been up to now limited both by the redshift range and by the relatively small sizes of the homogeneously surveyed areas.} {In order to enlarge publicly available optical cluster catalogs, in particular at high redshift, we have performed a systematic search for clusters of galaxies in the Canada France Hawaii Telescope Legacy Survey (CFHTLS).} {We considered the deep 2, 3 and 4 CFHTLS Deep fields (each 1$\\times$1~deg$^2$), as well as the wide 1, 3 and 4 CFHTLS Wide fields. We used the Le Phare photometric redshifts for the galaxies detected in these fields with magnitude limits of i'=25 and 23 for the Deep and Wide fields respectively. We then constructed galaxy density maps in photometric redshift bins of 0.1 based on an adaptive kernel technique and detected structures with SExtractor at various detection levels. In order to assess the validity of our cluster detection rates, we applied a similar procedure to galaxies in Millennium simulations. We measured the correlation function of our cluster candidates. We analyzed large scale properties and substructures, including filaments, by applying a minimal spanning tree algorithm both to our data and to the Millennium simulations. } {We have detected 1200 candidate clusters with various masses (minimal masses between 1.0 10$^{13}$ and 5.5 10$^{13}$ and mean masses between 1.3 10$^{14}$ and 12.6 10$^{14}$ M$_\\odot$) in the CFHTLS Deep and Wide fields, thus notably increasing the number of known high redshift cluster candidates. We found a correlation function for these objects comparable to that obtained for high redshift cluster surveys. We also show that the CFHTLS deep survey is able to trace the large scale structure of the universe up to z$\\geq$1. Our detections are fully consistent with those made in various CFHTLS analyses with other methods. We now need accurate mass determinations of these structures to constrain cosmological parameters.} {We have shown that a search for galaxy clusters based on density maps built from galaxy catalogs in photometric redshift bins is successful and gives results comparable to or better than those obtained with other methods. By applying this technique to the CFHTLS survey we have increased the number of known optical high redshift cluster candidates by a large factor, an important step towards using cluster counts to measure cosmological parameters.} ", "introduction": "\\label{sec:intro} The beginning of the 21st century is an exciting period for cosmological studies. Several methods now allow to put strong constraints on cosmological parameters. We can for example reconstruct Hubble diagrams (supernovae or tomography) or use directly the primordial fluctuation spectrum. In addition, the cluster count technique is probably the oldest one (see e.g. Gioia et al., 1990). Up to now this technique was penalized by the redshift range of detected clusters, which was too low to make the difference between flat and open universes. Distant cluster surveys have also been mainly conducted in areas too small or with inhomogeneous selection functions. Besides cluster mass knowledge, this technique requires indeed large fields of view of several dozen square degrees to provide large numbers of cluster detections at z$\\geq$1 (e.g. Romer et al. 2001). Recent X-ray cluster surveys are beginning to produce cluster catalogs at high z (e.g. the XMM-LSS survey, Pierre et al., 2007) and it is the goal of the present paper to contribute to the production of similar large cluster catalogs based on optical Canada France Hawaii Telescope Legacy Survey data. The Canada-France-Hawaii Telescope Legacy Deep and Wide Surveys (CFHTLS-D and CFHTLS-W) respectively explore solid angles of 4 deg$^2$ and 171 deg$^2$ of the deep Universe, each in 4 independent patches (http://www.cfht.hawaii.edu/Science/CFHLS/). For both surveys, observations are carried out in five filters ($u*,g',r',i'$ and $z'$) providing catalogs of sources that are 80\\% complete up to $i_{AB}$=26.0 (CFHTLS-D) and $i_{AB}$=24.0 (CFHTLS-W) (Mellier et al 2008, http://terapix.iap.fr/cplt/oldSite/Descart/CFHTLS-T0005-Release.pdf). The CFHTLS-W, in particular, encloses a sample of about 20 10$^6$ galaxies inside a volume size of $\\sim 1$ Gpc$^3$, with a median redshift of z$ \\sim 0.92$ (Coupon et al 2009). According to the standard cosmological model, the CFHTLS-W (W1, W2, W3, and W4 herafter) is then expected to contain 1000 to 5000 clusters of galaxies with accurate photometric redshifts, most of them in the $0.63$ as Miniutti et al. (2007) advocates. Difference is mainly due to the fact that we included warm absorbers, and used neutral refection without relativistic smearing. So that the constancy of the reflection component is explained by the light-bending model, all the disk reflection {\\em must}\\/ take place within 3 $r_g$, in which case the reflection normalization requires $\\Omega/2\\pi \\sim 3$ (Miniutti \\& Fabian 2004). On the other hand, our result favors a smaller solid angle of the reflector. Also, invariability of the neutral reflection component suggests that the reflection takes place far enough from the black hole, so that the intrinsic variation is smeared. If that is true, we would expect a narrow line of the equivalent width $ \\sim 150$ eV (e.g., George and Fabian 1991) for $\\Omega/2\\pi \\sim 1$, much larger than the observed value ($\\sim$18 eV; Young et al.\\ 2005). If we allow the line width free, we obtain the equivalent width $\\sim$ 110 eV (table \\ref{table0}), which may be reconciled with the reflector having $\\Omega/2\\pi \\sim$ 1. In this case, we require a mechanism to mildly broaden the line width up to $1 \\sigma \\approx 290$ eV. If this intrinsic line width is from Keplarian motion, with $v/c \\sim 0.29/6.35$, $v$ is $\\sim 14,000$ km/s, which does not seem too infeasible to be material on the innermost edge of the broad line region. In summary, we found a characteristic spectral variation of MCG-6-30-15, that is explained by variation of ionization degree of the warm absorber accompanying the flux variation, while the neutral reflection component was invariable throughout the observation. Considering the warm absorber, the observed energy spectrum is explained by a mildly broadened iron line, which is not significantly red-shifted, and a neutral and constant disk reflection component of which solid angle is $\\Omega/2\\pi \\sim 1$. Still, there needs a mechanism to keep the reflection component constant while direct component is variable." }, "0910/0910.5821_arXiv.txt": { "abstract": "In this letter, we report on dual-frequency European VLBI Network (EVN) observations of the faintest and least luminous radio cores in Seyfert nuclei, going to sub-mJy flux densities and radio luminosities around $10^{19}\\,$W\\,Hz$^{-1}$. We detect radio emission from the nuclear region of four galaxies (NGC\\,4051, NGC\\,4388, NGC\\,4501, and NGC\\,5033), while one (NGC\\,5273) is undetected at the level of $\\sim 100\\, \\mu$Jy. The detected compact nuclei have rather different radio properties: spectral indices range from steep ($\\alpha>0.7$) to slightly inverted ($\\alpha = -0.1$), brightness temperatures vary from $T_B=10^5$ K to larger than $10^7$ K and cores are either extended or unresolved, in one case accompanied by lobe-like features (NGC\\,4051). In this sense, diverse underlying physical mechanisms can be at work in these objects: jet-base or outflow solutions are the most natural explanations in several cases; in the case of the undetected NGC\\,5273 nucleus, the presence of an advection-dominated accretion flow (ADAF) is consistent with the radio luminosity upper limit. ", "introduction": "Active Galactic Nuclei (AGN) are traditionally divided in radio quiet (RQ) and radio loud (RL). The latter are typically powerful radio sources ($P_r>10^{22}\\,$W\\,Hz$^{-1}$), with large scale radio lobes and bright compact cores; VLBI observations routinely target their nuclear regions, showing high brightness temperatures and in some cases jet knots with superluminal motions. Radio quiet AGN, such as Seyfert galaxies, are much fainter radio sources and their radio emission is confined to the sub-kpc scale. However, moderately deep VLA surveys show that most AGN are radio sources at some level \\citep[][~hereinafter HU01]{Nagar2002,Ho2001}. While the origin of the radio emission in RL AGN is well established as synchrotron radiation from energetic particles in jets and lobes, the case for RQ AGN is much less clear. Since nuclear structures in most Seyfert galaxies show complex morphologies, it is of fundamental importance to resolve them with high spatial resolution. Indeed, it has been shown that VLBI observations of the pc-scale and subpc-scale region of Seyfert nuclei and Low-luminosity AGNs (LLAGNs) are often successful in the determination of the physical parameters of the nuclear radio components (such as the brightness temperature, the spectral index, the jet motions, etc.) and therefore in the comprehension of the underlying physical mechanisms. For example, the VLBA study of the classical Seyfert 2 galaxy NGC\\,1068 has allowed the identification of the location of the hidden active nucleus and the attribution of the core radio emission to thermal free-free emission from an X-ray heated corona or wind arising from the disk \\citep{Gallimore2004}. In the case of the type 1 Seyfert NGC~4151 it has been possible to resolve the 0.2 pc two-sided base of a jet whose low speeds indicate non-relativistic jet motions, possibly due to thermal plasma \\citep{Ulvestad2005}. On the contrary, the VLBA analysis of the LLAGN NGC\\,4278 has shown that the radio emission of this source is emitted via synchrotron process by relativistic particles similarly to ordinary radio-loud AGN \\citep{Giroletti2005}. These studies indicate that the analysis of individual sources is very important for the determination of the different spatial components and physical parameters. The five sources here presented belong to a complete distance limited ($d < 22\\,$Mpc) sample of 27 nearby Seyfert galaxies \\citep{Cappi2006}. Accurate multi-wavelength studies are available for this sample. In particular, VLA radio images at 1.4 and 5 GHz are presented in \\citetalias{Ho2001}, and VLA data at 15 GHz are also available for $25$ sample sources \\citep{Nagar2002}. VLBI observations have become available for a few sources through the years; $9$ sources of the sample have 5 GHz VLBA observations \\citep{Nagar2002}, while for the radio brightest galaxies, e.g.\\ NGC\\,1068 and NGC\\,3031, dedicated works are available \\citep{Gallimore2004,Bietenholz2004}. However, most of the weakest sources in the sample have never been observed with milliarcsecond resolution. In the present letter, we report on new high sensitivity VLBI observations of five sources with $S\\sim1$ mJy, aimed at answering two fundamental questions: (1) how common/frequent is the presence of (sub-)parsec scale radio sources even in the weakest nuclei and (2) what do the properties of the detected radio sources tell us about the physics of individual sources and the viability of jet-base versus Advection-Dominated Accretion Flow \\citep[ADAF,][]{Narayan1994} explanations. The observations are detailed in \\S2 and the results presented in \\S3. A discussion and the main conclusions are given in \\S4 and \\S5, respectively. Throughout the paper, we define spectral indices such that $S(\\nu)\\propto \\nu^{-\\alpha}$ and adopt $H_0=70$ km s$^{-1}$ Mpc$^{-1}$, $\\Omega_\\Lambda=0.73$ and $\\Omega_m=0.27$ \\citep{Spergel2003}. ", "conclusions": "With the VLBI observations presented in this letter, we have targeted the faintest and least luminous nuclei among well known local AGN, going to sub-mJy flux densities and radio luminosities around $10^{19}$ W Hz$^{-1}$. These sources belong to a complete sample of nearby Seyfert galaxies, in which previous VLBI observations -- albeit limited to the brightest members -- had revealed an ubiquitous presence of sub-parsec cores and/or structures. Similar findings had been reported on other relatively bright flat spectrum radio sources in LLAGN; for instance, \\citet{Nagar2002} found that almost all LLAGNs in the Palomar sample with S$_\\mathrm{VLA, 15\\, GHz}> 2.7$ mJy show mas-scale or sub mas-scale radio cores. The 80\\% detection rate in the present sub-sample extends these findings to the lowest luminosity and flux density regimes. While the presence of sub-pc scale radio emission appears thus to be ubiquitous, it is remarkable that it accounts only for a fraction between 5\\% and 40\\% of the emission detected on scales of a few tens of parsecs (eg.\\ as seen in VLA observations). As a consequence, a large fraction of the parsec scale radio luminosity is emitted in a diffuse region. This could be the case of NGC\\,4388 and NGC\\,4501, in which we completely resolve the VLA sources at 5 GHz, but detect them at 1.6 GHz (size $\\ga 10$ mas). Similarly, the non detection of NGC\\,5273 requires that $>95\\%$ of the VLA 1.6 GHz flux density is emitted on scales larger than 20 mas (1.6 pc). Even though much of the intermediate scale emission is resolved out, the majority of the nuclear regions do host compact radio sources at sub-parsec scales. In Table~\\ref{tabtot}, we report the multi-wavelength properties of the sources presented in this letter. The common features in these nuclei are: (i) their extremely low radio luminosity (Cols.\\ 4-6), at the level of the least luminous Seyfert nuclei \\citep[such as the sub-mJy source NGC\\,4395,][]{Wrobel2006}; (ii) their extreme radio-quietness, for instance the ratios between the radio and nuclear X-ray emission (Col.\\ 8) are among the lowest ever measured for LLAGN \\citep{Panessa2007,Terashima2003}; and (iii) their low Eddington ratios (Col.\\ 9), a common trait in low luminosity nuclei \\citep{Ho2009}. Radio emission from LLAGN cores is generally associated with accretion/ejection processes in the vicinity of a supermassive black hole \\citep{Falcke2000,Ulvestad2001,Nagar2002}. While some of the resolved extended emission could be of thermal origin, the small linear scales ($< 0.6$ pc) and comparatively high brightness temperatures ($T_B=10^{5-7}$ K) measured in our sources suggests that the weakest Seyfert nuclei could also be scaled-down versions of more luminous AGN. In this scenario, the origin of radio emission can be generally attributed to synchrotron emission from the base of a jet coupled with a low-power accretion disk \\citep{Falcke1999}. Alternatively, the presence of an ADAF \\citep{Narayan1994} is invoked to explain the low luminosity of these sources, although the predicted radio emission often fails to reproduce the data, requiring combined jet/ADAF models \\citep[see e.g.,][]{Yuan2002}. On average, our results do not seem consistent with the presence of an ADAF alone, on the basis of the observed structures, sizes, flux densities, and spectral indices. The two sources detected at both 1.6 and 5 GHz (NGC\\,4051 and NGC\\,5033) are most easily explained in terms of jet-base/outflow phenomena. First, the detection of three aligned sub-mJy components in NGC\\,4051 is indeed suggestive of ejection processes. The brightness temperature ($T = 2\\times10^5$ K) does not completely rule out a thermal origin, consistent with the presence of an outflow rather than a relativistic jet, similar to the case of NGC\\,4395, which shows an elongated low brightness temperature nuclear structure, tracing possible outflow emission \\citep{Wrobel2006,Christopoulou1997}. The EVN radio core is also positionally coincident with a low luminosity H$_2$O maser, suggesting that the radio continuum may arise from the inner regions of a molecular disk or from a nuclear wind \\citep{Hagiwara2007}. As for NGC\\,5033, its core presents all the hallmarks of a jet-base feature, being detected and unresolved at both at 1.6 and 5 GHz, with a brightness temperature lower limit of $T_B>1.3\\times10^7$ K, and a flat spectrum (consistent with $\\alpha=0$). While these would somewhat be consistent with ADAF expectations, the 5 GHz luminosity is $\\sim$ 4 times in excess of what predicted on the basis of the observed L$_\\mathrm{2-10 keV}$ luminosity, implying that the ADAF model alone fails to account for the observed radio emission. The presence of a jet or an outflow component is therefore required and more consistent with the data. Indeed, the short jet-like extension found at larger scale with VLA \\citepalias{Ho2001,Perez-Torres2007} favours the jet-base hypothesis. With all the caveats related to the unfavourable observing conditions at 5 GHz, the two sources detected only at 1.6 GHz (NGC\\,4388 and NGC\\,4501) seem to be also at odds with an ADAF scenario. The non detection at high frequency points to a steep spectral index and/or to a rather extended structure, at scales of $10^6 R_S$ or more, in contrast with the flat/inverted spectrum and the compact structure predicted by ADAF models. Although consistent with an angular scale of several mas, thermal emission is also unlikely given the $\\ga 10^6$ K brightness temperatures observed at 1.6 GHz. NGC\\,4388 and NGC\\,4501 are also the only type 1.9 Seyferts in our sample, showing heavily absorbed/weak X-ray spectra \\citep{Cappi2006}. Interestingly, the only undetected source in our sample is NGC\\,5273, a type 1.5 Seyfert, in which the nucleus is seen directly, displaying broad emission lines and a bright X-ray spectrum \\citep{Cappi2006}. Either the VLA emission is resolved out or the source is variable. Indeed, the source was initially not revealed at 8 GHz with the VLA \\citep[$S < 0.23$ mJy,][]{Kukula1995} and lately detected with $S = 0.6$ mJy by \\citet{Nagar1999}. The comparison between the X-ray and the EVN radio luminosity upper limit reveals that this source is extremely radio quiet. Indeed, this is the only case in which the radio data are consistent with a pure ADAF accretion mechanism, since our upper limit for this nucleus is 4 times above the radio core luminosity derived from the observed X-rays \\citep{Yi1998}." }, "0910/0910.1069_arXiv.txt": { "abstract": "From multi-epoch adaptive optics imaging and integral field unit spectroscopy we report the discovery of an expanding and narrowly confined bipolar shell surrounding the helium nova V445 Puppis (Nova Puppis 2000). An equatorial dust disc obscures the nova remnant, and the outflow is characterised by a large polar outflow velocity of 6720 $\\pm$ 650 km s$^{-1}$ and knots moving at even larger velocities of 8450 $\\pm$ 570 km s$^{-1}$. We derive an expansion parallax distance of 8.2 $\\pm$ 0.5 kpc and deduce a pre-outburst luminosity of the underlying binary of $\\log L/L_{\\odot} = 4.34 \\pm 0.36$. The derived luminosity suggests that V445 Puppis probably contains a massive white dwarf accreting at high rate from a helium star companion making it part of a population of binary stars that potentially lead to supernova Ia explosions due to accumulation of helium-rich material on the surface of a massive white dwarf. ", "introduction": "Accretion onto compact objects can lead to a range of explosive phenomena since their accreted layers can reach densities and temperatures high enough to initiate nuclear burning. Helium novae are expected to occur during periods of high mass-accretion rates ($\\dot{M} \\sim 10^{-9} - 10^{-6}$ M$_{\\odot}$ yr$^{-1}$) of He-rich material through unstable helium burning via helium shell flashes \\citep{iben91}. In such events, typically $\\sim 10^{-4}- 10^{-2}$ M$_{\\odot}$ of fuel is ignited on the surface of the accretor \\citep{kato99,yoon04,bild07}. Models of the binaries that experience such helium novae generally consist of a 0.6 -- 0.8 M$_{\\odot}$ CO white dwarf (the accretor) with a lower-mass H-deficient companion (the donor). The donor can be either a non-degenerate helium star \\citep[e.g.,][]{yoon04} or a semi-/fully degenerate star. The latter are usually referred to as AM CVn systems, representing a class of ultra-compact helium-transferring binaries \\citep{nele04,bild07,roel07}. Some of these compact binaries may lead to sufficient mass accumulation onto the primary to be viable supernova Ia progenitors after successive He novae \\citep{iben94,kato89,yoon03,tutu07,wang09}. Accreting massive white dwarfs in close binary systems are the favoured progenitors of supernova Ia explosions \\citep{branch07,par07}. Possible binary channels are typically divided into single-degenerate scenarios involving hydrogen-rich companions versus double-degenerate progenitors. The latter include the white dwarf plus He-star and double white dwarf pathways. To date, only about a dozen promising single-degenerate type Ia progenitors have been identified \\citep[see, e.g.,][]{par07}. U~Scorpii and RS Ophiuchi represent the two different classes of hydrogen-rich companion stars in the supernova Ia progenitors: main-sequence or slightly evolved stars, and red giants, respectively \\citep{li97}. Ongoing supernova searches are, however, revealing a growing diversity of type Ia explosion events, which challenge current formation scenarios. It could be that, despite their low predicted frequency, the helium star donor channel to type Ia explosion is successful in explaining some of the diversity observed, such as the population of young Ia progenitors \\citep{manu06,wang09}. \\object{V445 Puppis} (Nova Puppis 2000) is the first, and so far only, helium nova detected \\citep{ashok03,kato03}. The nova outburst of V445 Puppis was first noticed on 23 November 2000 \\citep{kato01}, and the optical outburst characteristics are those of a slow nova. Optical \\citep{wagn01,iijim08}, near-infrared \\citep{ashok03} and mid-infrared \\citep{lynch01} spectra of V445 Puppis obtained during outburst immediately revealed its most peculiar and defining characteristic: the complete absence of hydrogen in the ejecta. The outburst light curve has been modelled \\citep{kato03,kato08} by free-free emission and an optically thick wind, following a helium shell flash on the surface of an accreting massive white dwarf. A lower limit on the mass of the accreting white dwarf of $M$ $\\ge$ 1.35 M$_{\\odot}$ is inferred by \\citet{kato08}, making V445 Puppis a possible progenitor of a supernova Ia from a helium-rich donor channel. In these models, the white dwarf is expected to grow in mass as the mass ejected through repeated helium nova outbursts is less than the accreted mass \\citep{kato99}. V445 Puppis provides the first empirical benchmark for a helium-dominated outburst on the surface of a white dwarf. In Section 2 we present new observations of V445 Puppis taken 5 -- 7 years after outburst, and determine the distinct evolution of the resolved nova shell from a spatio-kinematic analysis (Section 3). This leads to a robust distance determination to V445 Puppis. In Section 4, we discuss the implications of the distance derived here on the nature of the underlying binary. ", "conclusions": "\\subsection{Pre-outburst conditions} With the distance known, the nature of the helium nova progenitor can be constrained from pre-outburst observations. Unfortunately, not much is known of V445 Puppis prior to outburst. Apart from optical ($V = 14.5$ mag, from VSNET, see \\citet{ashok03}) and near-infrared (2MASS: $K_s = 11.52$ mag) flux measurements (see lower panel of Fig.~\\ref{woudtfig1}), neither spectrum nor orbital period have been obtained prior to outburst. We verified the plate archives at the Harvard-Smithsonian Center for Astrophysics for prior outbursts (or sustained long periods of obscuration indicative of a missed outburst). Despite good time coverage and the fact that V445 Puppis could be identified on many plates at approximately constant brightness (based on visual comparision with nearby stars in the field), we could find no signatures of a previous outburst in the 1897 -- 1955 time frame. V445 Puppis is located at the low Galactic latitude of $b = -2.19^{\\circ}$, which at a distance of 8.2 $\\pm$ 0.5 kpc translates to a height below the Galactic Plane of 313 $\\pm$ 19 pc. At that distance, and that far below the Galactic Plane, one can use the IRAS/DIRBE Galactic reddening maps \\citep{schleg98} to gauge the Galactic reddening towards V445 Puppis. In Fig.~\\ref{woudtfig6} we show the ratio of the measured reddening (at a given distance) to the total line-of-sight Galactic reddening for open clusters in the Milky Way \\citep{khar05} in the Galactic latitude range $1^{\\circ} \\le |b| \\le 4^{\\circ}$ and for $E(B-V)_{\\rm Schlegel} \\le 2.5$, as a function of height above/below the Plane and distance, respectively. The reddening maps of \\citet{schleg98} are not well-calibrated at low Galactic latitude; from colors of galaxies discovered at low Galactic latitude behind the southern Milky Way it appears that the reddening values from Schlegel et al.~are too high \\citep{schroed07} and that the true value is 87\\% of the \\citet{schleg98} reddening, e.g. $E(B-V)^{\\rm cal}_{\\rm Schlegel} = 0.87 E(B-V)_{\\rm Schlegel}$. Comparing the location of V445 Puppis to open clusters within 10 degrees of V445 Puppis (big filled circles in Fig.~\\ref{woudtfig6}), we deduce that Galactic foreground reddening towards V445 Puppis is probably in the range of $E(B-V) = 0.51 - 0.68 = 0.75 - 1.0 \\, E(B-V)^{\\rm cal}_{\\rm Schlegel}$, as marked by the arrow in Fig.~\\ref{woudtfig6}. The lower limit given here could be a conservative one due to a selection effect; given the relative poor angular resolution of the reddening maps, the open clusters which set the lower ratio limits (those observed at large distances and large height above/below the Galactic plane) are likely to be observed through small-scale patches of relatively lower extinction compared to their surrounding. Finally, taking the equivalent widths of the two interstellar Na D components \\citep{iijim08}, a value of $E(B-V) = 0.62$ is inferred using the calibration of \\citet{mun97}. This value falls within our deduced limits. Taking the lower limit of the Galactic reddening towards V445 Puppis ($E(B-V) = 0.51$), we derive a pre-outburst color of V445 Puppis of $(V-K)^0 = 1.58$ mag, assuming a standard Galactic extinction law \\citep{card89}. This is significantly redder than the likely binary progenitors -- high $\\dot{M}$ AM CVn systems or systems with helium stars donors -- which typically have $(V-K)^0 \\approx -0.6$. A red giant donor (symbiotic nova) can be excluded on the basis of the pre-outburst colors. This suggests the presence of substantial circumstellar (CS) reddening before the November 2000 outburst, which is not unreasonable given the possibility of previous outbursts or material expelled during a common envelope phase. The CS reddening around V445 Puppis does not necessarily have to follow a standard Galactic extinction given the current carbon-rich ejecta; \\citet{berge99} find some deviations from a standard reddening law for CS reddening around a number of carbon-rich R CrB stars. In the right panel of Fig.~\\ref{woudtfig3}, we compare the line profile of the 7065 {\\AA} and 2.0581 $\\mu$m He\\,I recombination lines, obtained close in time and both normalised by peak blueshifted emission. As expected, the redshifted component of the optical line (at +1140 km s$^{-1}$) appears dimmer compared to its near-infrared counterpart, due to dust obscuration within the shell. For a standard Galactic reddening law \\citep{card89}, we expect a ratio of peak emission of near-infrared (2.0581 $\\mu$m) to optical (7065 {\\AA}) of 7. The ratio observed in V445 Puppis is 3.4, substantially less and indicative of a low ratio of total-to-selective extinction, $R_V \\approx 2.5$ \\citep{fitzp99}. The shell of V445 Puppis offers an opportunity to determine the nature of the dust extinction in carbon-rich outflows. The new X-shooter instrument on the VLT will be ideal to obtain simultaneous optical to near-infrared medium-to-high resolution spectroscopy of both NE and SW shells. To make the pre-outburst color of V445 Puppis consistent with the intrinsic colors of likely progenitor binaries, a (maximum) additional color excess of $A_V - A_K$ = 2.2 mag requires $A_V$ = 2.5 mag for $R_V = 3.1$ (standard reddening law), or $A_V$ = 2.8 mag for $R_V = 2.5$. The uncertainty in $A_V$ introduced by the different reddening laws, however, is small compared to the uncertainty in the bolometric correction needed to arrive at the pre-outburst luminosity of V445 Puppis. We thus determine a pre-outburst extinction-corrected brightness of at least $V^0 = 12.9$ mag (corrected for Galactic reddening only), but more likely $V^{0} \\approx 10.1 - 10.4$ mag (including circumstellar reddening correction), with the range in the latter allowing for differences in the CS reddening law. The corresponding absolute magnitudes are $M_V^{0} = -1.7$ and $-4.2$ to $-4.5$ mag, respectively. For a range of likely bolometric corrections (BC = $-1$ to $-2.5$, \\citet{roel07,heb81}), the pre-outburst luminosity of V445 Puppis is $\\log L/L_{\\odot} = 3.3 \\pm 0.3$ (corrected for Galactic extinction only), or $\\log L/L_{\\odot} = 4.34 \\pm 0.36$ (including circumstellar reddening correction). It has not escaped our attention that hydrogen-deficient carbon (HDC) stars also occupy this luminosity regime. With a larger distance -- this paper -- and a larger Galactic reddening correction \\citep{iijim08}, the possibility that V445 Puppis belongs to the class of HDC stars can no longer be rejected on luminosity grounds, cf. \\citet{ashok03}. The (Galactic) extinction-corrected $(V-K_S)^0$ colors of V445 Puppis are not too dissimilar from the HDC star HD\\,182040 \\citep{brun98}. It is not clear, though, what can cause a luminosity increase of 6 magnitudes and a rapidly expanding shell from a HDC star; the overall light curve appears much better described by an event near a compact, massive white dwarf. \\subsection{Shaping of the nebula} Deviations from the simple outflow model presented in Section 3 could provide clues to the mechanism responsible for shaping the nebula. Although the dynamical model fits the shell remarkably well, the largest deviations occur nearest to the nova remnant at the earliest epochs. The nebula initially had a very narrow waist, followed by a rapid broadening of the waist close to the nova remnant. To produce very narrow waists in PNe, collimated fast winds (CFW) are needed in systems with high density gas in an equatorial plane close to the source of the CFW \\citep{sok00}. It is unclear what collimated the fast wind in V445 Puppis. At the moment it seems likely that an equatorial disk/torus already existed pre-outburst, given the large pre-outburst circumstellar reddening deduced in Section 4.1. Once the present optically-thick dust disk clears (Fig.~\\ref{woudtfig2} -- presumably additional material was ejected in the equatorial plane during the November 2000 outburst) and a relatively unobscured view of the nova remnant is possible, alternative origins of the collimated outflow could be investigated, e.g. the presence of a rapidly rotating magnetic white dwarf. The shell of V445 Puppis is unique for a classical nova. Although many novae show bipolar shells, for example the classical nova HR Del \\citep{har03} and recurrent nova RS Oph \\citep{bode07,sok08}, the shell in V445 Puppis reflects the most collimated outflow seen in any nova. The knots are moving at jet-like velocities of $\\ga$ 8000 km s$^{-1}$. They can be a consequence of a jet-activity about 350 days after the main outburst which is not long after the sudden drop off in the $V$ band is seen \\citep{ashok03} and coincides with the radio flare \\citep{rup01}. In the symbiotic nova CH Cygni, a drop in $V$ band magnitude is associated with the onset of a radio jet ejection, see e.g. \\citet{karov07}. \\subsection{V445 Puppis as a proto-type helium nova} At a distance of 8.2 kpc, the maximum brightness at outburst ($V = 8.6$ mag) is consistent with the Eddington luminosity of a massive white dwarf. Moreover, the pre-outburst luminosity of V445 Puppis of $\\log L/L_{\\odot} = 4.34 \\pm 0.36$ appears to rule out an AM CVn progenitor and may require both a luminous accretion flow as well as a bright donor. For comparison, a 1.3-M$_{\\odot}$ white dwarf accreting at $\\dot{M} \\sim 10^{-6}$ M$_{\\odot}$ yr$^{-1}$ would give an expected accretion luminosity of $\\log L/L_{\\odot} = 3.7$. Similarly, the luminosity of a moderately massive, evolved helium star is around $\\log L/L_{\\odot} \\approx 3.7$ \\citep{yoon03}. With the distance determined in this paper, a revised mass of the dust shell of $1.5 \\times 10^{-5}$ M$_{\\odot}$ is derived following \\citet{lynch04}. This is consistent with \\citet{kato03}, but we note that the mass depends strongly on the assumed temperature of the shell and will ultimately be best constrained by mid- to far-infrared data of V445 Puppis recently obtained by {\\it Spitzer}. This mass is likely to be a lower limit given the large optical depth of the equatorial dust disk. The various components of V445 Puppis (the dust disk, the bipolar shell) clearly show that the immediate environment of V445 Puppis is sculpted either directly or indirectly by (repeated) outbursts. If V445 Puppis is representative of its class (of helium novae), the helium counterparts to the classical hydrogen-rich novae result in more substantial circumbinary reddening due to the carbon-rich outflow. Validation of the white dwarf + helium star model as the appropriate binary configuration of V445 Puppis will come from the determination of its orbital period. Unfortunately, this can only be done once the optically thick dust disk surrounding the nova remnant becomes transparent. Given the low inclination of the bipolar outflow, the implied inclination of the equatorial plane which corresponds to the orbital plane of the binary is $\\sim 86^{\\circ}$. We thus expect this to be an eclipsing binary, which would further facilitate our ability to constrain the component masses. \\subsection{V445 Puppis as a candidate Ia progenitor} There are several indirect suggestions that the white dwarf in V445 Puppis is massive: the maximum luminosity during outburst, the large velocities of the ejected blobs, and its pre-outburst (accretion) luminosity. The lack of strong soft X-ray emission pre-outburst, as is expected from a supersoft source \\citep{yoon03} could be explained by the substantial interstellar and circumstellar absorption which currently totally obscures the underlying accreting binary. V445 Puppis has not been detected in the ROSAT all-sky survey. Our derived luminosity is consistent with the massive white dwarf + helium star model of \\citet{yoon03}, but see also \\citet{kato08}. It is likely that V445 Puppis belongs to a class of binaries that, in principle, could lead towards supernovae of type Ia. Given that the expected lifetime of a massive white dwarf + helium star binary is short, these progenitors are associated with the relatively young ($\\sim 10^8$ yr) population of Ia SNe. Whether V445 Puppis itself eventually leads to a supernova Ia event, or if the current helium nova has pre-empted that pathway, depends critically on the mass accumulation efficiency on the surface of the white dwarf and the mass ejected during this nova outburst. This remains a topic of debate." }, "0910/0910.4961_arXiv.txt": { "abstract": "As the Galaxy evolves, the abundance of deuterium in the interstellar medium (ISM) decreases from its primordial value: deuterium is ``astrated\". The deuterium astration factor, $f_{\\rm D}$, the ratio of the primordial D abundance (the D to H ratio by number) to the ISM D abundance, is determined by the competition between stellar destruction and infall, providing a constraint on models of the chemical evolution of the Galaxy. Although conventional wisdom suggests that the local ISM (i.e., within $\\sim 1-2$~kpc of the Sun) should be well mixed and homogenized on timescales short compared to the chemical evolution timescale, the data reveal gas phase variations in the deuterium, iron, and other metal abundances as large as factors of $\\sim 4-5$ or more, complicating the estimate of the ``true\" ISM D abundance and of the deuterium astration factor. Here, assuming that the variations in the observationally inferred ISM D abundances result entirely from the depletion of D onto dust, rather than from unmixed accretion of nearly primordial material, a model-independent, Bayesian approach is used to determine the undepleted abundance of deuterium in the ISM (or, a lower limit to it). We find the best estimate for the undepleted, ISM deuterium abundance to be (D/H)$_{\\rm ISM}\\geq (2.0 \\pm 0.1) \\times 10^{-5}$. This result is used to provide an estimate of (or, an upper bound to) the deuterium astration factor, $f_{\\rm D} \\equiv$~(D/H)$_{\\rm P}/$(D/H)$_{\\rm ISM} \\leq 1.4 \\pm 0.1$. ", "introduction": "Deuterium is created in an astrophysically interesting abundance only during big bang nucleosynthesis (BBN) \\citep{boes,steigman07}, after which, in the post-BBN Universe, its abundance ($y_{\\rm D} \\equiv 10^5 (\\rm D/H)$) decreases monotonically due to the processing of gas through succeeding generations of stars where deuterium is completely destroyed \\citep{els,pf03}. Consequently, deuterium plays a special role in cosmology, nuclear astrophysics, and in Galactic chemical evolution \\citep{ytsso,boes,st92,st95,vangioni,tosi96,schramm,romano06,srt,steigman07,pf08}. Its relatively simple Galactic evolution permits us to use deuterium to determine the fraction of interstellar gas that has been processed through stars \\citep{st92,st95}. By comparing the primordial and Galactic deuterium abundances one can learn about and discriminate among different Galactic chemical evolution models \\citep{vangioni,tosi96,romano06,srt,pf08}. Together with non-BBN constraints on the baryon (nucleon) density from observations of the cosmic microwave background \\citep{spergel,Dunkley09,komatsu09,komatsu10}, the primordial deuterium abundance, \\ydp, is predicted by BBN \\citep{cyburt03,coc,steigman07,vs,cyburt08}. The predicted abundance is in excellent agreement with observations of deuterium in high-redshift, low-metallicity QSO absorption line systems (QSOALS) \\citep{omeara,pettini}. Since deuterium is destroyed in the Galaxy as gas is cycled through stars, (D/H)$_{\\rm ISM} \\leq$ (D/H)$_{\\rm P}$. However, all successful chemical evolution models require infall to the disk of the Galaxy of unprocessed (or, nearly unprocessed) gas \\citep{tosi96,pf08}, and such deuterium-rich (and metal-poor) gas would raise the ISM D/H ratio closer to the primordial value. As a result, comparison of the primordial and ISM deuterium abundances provides a constraint on infall and, therefore, on chemical evolution models. Observations over the past decade and more of the deuterium abundance in the relatively local interstellar medium (ISM) reveal an unexpectedly large scatter in D/H \\citep{jenkins,sonneborn,hebrard,hoopes}, challenging the conventional wisdom of a well mixed ISM. For example, as shown in Figure~\\ref{fig:lgyvslgh}, the absorption line measurements from the {\\it Far Ultraviolet Spectroscopic Explorer} (FUSE) reveal variations of a factor of $\\sim 4$ ($0.5 \\la y_{\\rm D,ISM} \\equiv 10^{5}$(D/H)$_{\\rm ISM} \\la 2.2$) in the {\\em gas-phase} D/H ratios over lines of sight (LOS) to background stars within $\\sim 1-2$ kpc of the Sun. Moreover, the variations in the observed, gas phase D/H abundances are found to correlate {\\em positively} with the abundances of refractory elements such as Ti \\citep{prochaska,lallement08}, and a similar correlation is also found between D and other metals such as Fe and O \\citep{linsky,srt}. However, this positive correlation between the abundances of D and the metals is opposite to the trend expected from stellar nucleosynthesis (D decreasing as the metals increase). Motivated by the observed correlations and by the very large spread in the D/H ratios inferred from the FUSE and other data sets, it has been proposed that the large variations in {\\em gas-phase} D/H may be due to the depletion of gas phase deuterium onto dust grains \\citep{jura,draine04,draine06}. If this is the case, then the FUSE (and other) gas-phase absorption-line measurements reveal that depletion has not been well mixed (homogenized) in the ISM and, the data may only provide a {\\em lower} limit to the {\\em true} (undepleted) ISM deuterium abundance and, therefore, only an {\\em upper} limit to the deuterium astration factor, $f_{\\rm D} \\equiv y_{\\rm DP}/y_{\\rm D,ISM}$. \\begin{figure}[t] \\epsscale{0.4} \\plotone{lgyvslgh1.eps} \\caption{The logs of the deuterium abundances versus the logs of the \\hi~column densities [cm$^{-2}$] for the 49 FUSE LOS (see the text). The filled symbols are for the 41 LOS which have iron abundance data, while the open symbols are for the 8 LOS which lack iron abundances. The squares (blue) are for the LOS within the Local Bubble (LB) and the circles (red) are for the non-LB (nLB) LOS. The solid line is at the mean D abundance for the LB LOS (log $y_{\\rm D,LB} = 0.19$); the dashed line is its extension to the nLB LOS.} \\label{fig:lgyvslgh} \\end{figure} In Figure~\\ref{fig:lgyvslgh} the logs of the deuterium abundances (log~$y_{\\rm D} \\equiv 5~+~$logN(\\di) $-$ logN(\\hi)) are shown as a function of the logs of the \\hi~column densities for the 49 LOS with \\di~column densities from Table 2 of \\citet{linsky}, supplemented by data from \\cite{oliveira2006} and \\cite{dupuis2009}. Of these 49 LOS, 41 have iron abundance measurements (filled symbols); 21 of the 49 LOS are within the Local Bubble (LB; see \\citet{linsky} for a discussion of the LB); the remaining 28 non-Local Bubble (nLB) LOS are toward stars beyond the LB (see \\S2.1). The unexpectedly large spread among the observationally inferred ISM D abundances complicates any estimate of \\ydism. Recognizing this point, \\citet{linsky} chose for their estimate of a {\\em lower bound} to the true (undepleted) ISM deuterium abundance the mean of the five {\\em highest} D/H ratios finding \\ydism~$\\geq 2.17 \\pm 0.17$ or, when including corrections (see the discussion in \\S2.1) for N(\\hi) and N(\\di) for those lines of sight outside of the Local Bubble, \\ydism~$\\geq 2.31 \\pm 0.24$. These estimates are quite close to the {\\it lower bound} to the primordial abundance estimated from the QSOALS, $y_{\\rm DP} = 2.82^{+0.20}_{-0.19}$ \\citep{pettini}\\footnote{ Since this estimate of \\ydp~relies on only 7 high redshift, low metallicity LOS and, since the dispersion in deuterium abundances among them is unexpectedly large, some prefer to adopt a so-called ``WMAP D abundance\". WMAP does not observe deuterium. Rather, the WMAP-determined baryon density parameter may be used in a BBN code to {\\it predict} the relic D abundance. If the \\cite{komatsu10} estimate of the baryon density is adopted, the BBN predicted primordial D abundance is \\ydp~$= 2.5 \\pm 0.1$. However, the WMAP collaboration also provides an estimate of the effective number of neutrinos, \\citep{komatsu10} which, when used along with their baryon density estimate in a BBN code, leads to a {\\it different} predicted primordial D abundance \\ydp~$= 3.0 \\pm 0.4$. The difference between these two predictions reflects the difference between the standard model effective number of neutrinos expected and the WMAP observed value, which is within $\\sim 1.5\\sigma$ of expectations. So, which, if either, ``WMAP D abundance\" is preferred? Here, we compare observations to observations, not to model-dependent predictions and, we adopt the \\cite{pettini} value for \\ydp.}, suggesting a small upper bound to the D astration factor, $f_{\\rm D} \\leq 1.30^{+0.14}_ {-0.13}$ or, an even smaller value, $f_{\\rm D} \\leq 1.22 \\pm 0.15$ for the LB-corrected, nLB deuterium abundances. To account for such a high ISM deuterium abundance and such a small D astration factor, a very high infall rate of pristine material would be needed, challenging many Galactic chemical evolution models \\citep{romano06,pf08}. By limiting themselves to the five highest D abundances, \\citet{linsky} ignore the lower deuterium abundances along many more LOS which are consistent with them within the errors, potentially biasing their estimates of \\ydism~and of $f_{\\rm D}$. In an attempt to address this issue, \\citet{srt} (SRT), used the 18 highest D/H ratios from the FUSE data (see Table 3 of \\citet{linsky}), finding an ISM D abundance of $y_{\\rm D,ISM} = 1.88 \\pm 0.11$, corresponding to a D-astration factor $f_{\\rm D} \\leq 1.50^{+0.14}_{-0.13}$, consistent with at least some of the otherwise successful chemical evolution models discussed in SRT. In fact, the data in Table 2 of \\citet{linsky} reveal that there are 19 LOS with central values of log \\yd~$\\geq 0.20$. The weighted mean (along with the error in the mean) for these 19 D abundances is log $y_{\\rm D,19} = 0.26 \\pm 0.01$, corresponding to $y_{\\rm D,19} = 1.8 \\pm 0.1$. For these 19 LOS the reduced $\\chi^{2} = 0.85$, confirms that the weighted mean provides a good description of their D abundances. As more LOS with lower D abundances are added, the weighted mean decreases, but the reduced $\\chi^{2}$ increases, so that \\ydism~$\\ga 1.8 \\pm 0.1$ is likely a robust {\\it lower} bound to the ISM D abundance and, $f_{\\rm D} \\la 1.5 \\pm 0.1$ (log $f_{\\rm D} \\la 0.19 \\pm 0.03$) is a robust {\\it upper} bound to the deuterium astration factor. Surely, there must be a better way to find a reliable estimate of the maximum value of the deuterium abundance in the local ISM while accounting for the observational errors in the individual D abundance determinations. In \\S 2 we describe a Bayesian analysis designed to find the best estimate of the maximum, gas phase (undepleted), ISM deuterium abundance, \\ydmax, from data with non-negligible errors and we apply it to the FUSE data set. On the assumption that the spread in the observed D abundances is the result of incompletely homogenized dust depletion, \\ydism~$\\geq$ \\ydmax~and $f_{\\rm D} \\equiv y_{\\rm DP}/y_{\\rm D,ISM} \\leq y_{\\rm DP}/y_{\\rm D,max}$. Our results are summarized and our conclusions presented in \\S 3. As already noted, the dust depletion hypothesis suggests that there should be a correlation between the D and Fe abundances. Although this trend appears to be present at some level \\citep{linsky}, a deeper analysis suggests that the simplest interpretation of the depletion hypothesis may need to be modified~\\citep{srt}. For example, since D is likely attached to the mantles of dust grains \\citep{tielens}, while Fe is mostly contained in the cores of the grains, D will be removed from grains more easily than Fe. Thus, although strong shocks may destroy grains completely, weak shocks may only remove D from grains while Fe might stay locked within their cores \\citep{draine79}. This effect then may explain the scatter observed in the relation between the D and Fe depletions. Thus, in a follow-up paper in preparation, the consistency of the depletion hypothesis with the FUSE data for D and Fe (and O) is investigated, and some possible modifications to its simplest version are explored \\citep{sp}. ", "conclusions": "In the decades prior to the current era of precision cosmology the primordial abundance of deuterium provided the only quantitative cosmological baryometer~\\citep{boes}. Although, at present, the deuterium abundance is only measured along seven, high-redshift, low-metallicity LOS to background quasars~\\citep{omeara,pettini}, the inferred primordial D abundance, \\ydp~= $2.8 \\pm 0.2$, is in excellent agreement with the non-BBN inferred baryon density parameter~\\citep{steigman07}. In the post-BBN Universe deuterium is destroyed as gas is cycled through stars, so that comparing the abundance of deuterium in the ISM of the Galaxy with the primordial D abundance provides an estimate of the virgin fraction of the ISM (\\ie the amount of gas presently in the ISM which has never been cycled through stars), constraining models of Galactic chemical evolution~\\citep{st92,st95,srt,pf08}. According to conventional wisdom the deuterium-free, metal-enhanced products of stellar nucleosynthesis should be well-mixed in the local ISM. In contrast, the FUSE data on the abundances of deuterium and several metals (\\eg iron, oxygen, etc.) along LOS within $\\sim 1 - 2$ kpc of the Sun reveal a much different picture. The FUSE~\\citep{linsky} and earlier observations \\citep{jenkins,sonneborn,hebrard,hoopes,prochaska} reveal unexpectedly large gas phase variations in \\yd~(and in the abundances of iron, oxygen, etc.) within the local ISM, as shown for FUSE deuterium data in Figure~\\ref{fig:lgyvslgh}. It has been proposed that the large variations observed in the local ISM D abundances can be accounted for by preferential depletion of deuterium (relative to hydrogen) onto dust \\citep{jura,draine04,draine06}, although incompletely mixed infall of relatively unprocessed, deuterium-enhanced, metal-free material may have contributed to some of the observed variations~\\citep{st92,st95,srt}. The large variations among the ISM D abundances, along with observational errors and the possible contributions from dust depletion and infall, complicate using the D observations to provide a robust estimate of the ISM D abundance which, in combination with the primordial D value, can lead to a constraint on the deuterium astration factor, $f_{\\rm D}$. The key question is, given the data (with its errors), how to find the best estimate of the ``true\", undepleted, ISM D abundance? Here, to address this question, the limits to the true, undepleted, ISM D abundance were investigated employing a model-independent Bayesian statistical analysis similar to that used by \\citet{hogan} to infer the primordial helium abundance from a set of helium abundance observations. It was assumed, along with \\citet{linsky}, that the spread in the observed D abundances is the result of incompletely homogenized D depletion onto dust in the local ISM. In our analysis this is modeled by five different probability distributions (priors) for the \\yd~values. The \\yd~(actually, log~\\yd)~values shown in Figure \\ref{fig:lgyvslgh} suggest that, given the relatively large errors, a uniform (top-hat) distribution, favoring neither low-D nor high-D may be a good approximation to the data. To explore the sensitivity of our result to the choice of the prior, we first considered two asymmetric distributions -- a positive-bias prior favoring low depletion, and a negative-bias prior favoring large depletion, as well as two other priors -- an M-shaped distribution favoring both low and high depletion, and a complementary, $\\Lambda$-shaped distribution. Using the FUSE deuterium observations along all 49 LOS \\citep{linsky,oliveira2006,dupuis2009}, we found the likelihoods in the \\{\\ydmax,$w$\\} plane for the five choices of the Bayesian priors (see Figures~\\ref{fig:allallpriors} and \\ref{fig:lbnlballM}). For all priors, the Bayesian analysis of the full data set requires significant depletion (\\eg $w \\neq 0$ at greater than 99.9\\% confidence). Comparing the maximum likelihood values for the five different distributions, we find that the bimodal, M-shaped distribution provides the best fit to the observed data (see Table 1). However, it is important to notice that the shapes of the priors require an additional assumption in our analysis, so that the M-shaped distribution is the most model-dependent. \\footnote{The M-shaped prior favors both low and high levels of D depletion while strongly disfavoring intermediate depletion, suggesting that two competing processes may be at work: depletion onto dust and evaporation from dust, perhaps due to exposure to shocks. To fit the M-prior scenario, both processes would have to be efficient and rapid to account for the deficit of intermediate D abundances. The distribution of the presently available data (with its errors) is inconclusive and does not strongly favor any of the adopted prior distributions. When more data become available, the Bayesian approach presented here may be used to learn more about the mechanism of deuterium depletion onto dust.} In contrast, the top-hat prior is the least model-dependent, favoring all levels of depletion equally. Given our ignorance of the detailed depletion mechanisms responsible for the observed scatter in the gas phase ISM deuterium abundances, we prefer to adopt for our estimate of the undepleted, ISM deuterium abundance, the result of the simplest, top-hat prior, whose maximum likelihood is similar to that of the best-fitting M-distribution, \\beq y_{\\rm D,ISM} \\geq y_{\\rm D,max} = 2.0 \\pm 0.1 = 2.0(1 \\pm 0.05) \\eeq This value is our best estimate of the true ISM D abundance based on the available deuterium observations in the local ISM and is independent of any model-dependent assumptions about galactic chemical evolution. Combining our result with \\ydp~= $2.8 \\pm 0.2 = 2.8(1 \\pm 0.07)$ \\citep{pettini} (which, recall, provides a {\\it lower} bound to the primordial abundance), yields a limit to the deuterium astration factor $f_{\\rm D} \\leq 1.4 \\pm 0.1$ (for the M-prior, $f_{\\rm D} \\leq 1.6 \\pm 0.1$), consistent with most, but not all, Galactic chemical evolution models \\citep{srt,pf08,romano09}. If, on the other hand we compared this \\ydism~value to the BBN + WMAP inferred primordial D abundance, \\eg \\ydp~= $2.5 \\pm 0.1$ \\citep{steigman10}, \\ydp~= $2.5 \\pm 0.2$ \\citep{cyburt08} or, the prediction inferred when including the WMAP-determined effective number of neutrino species \\citep{komatsu10}, \\ydp~= $3.0 \\pm 0.4$, the resulting deuterium astration factor would be somewhat lower in first two cases, $f_{\\rm D} \\approx 1.3 \\pm 0.1$, which is marginally problematic for some GCE models. In contrast, for the \\citet{komatsu10} value of \\ydp, $f_{\\rm D} \\leq 1.5 \\pm 0.2$, which is entirely consistent with GCE models. As seen in Figures~\\ref{fig:lgyvslgh} -- \\ref{fig:lbnlball}, for the LB there is little scatter among the gas-phase D abundances. The small scatter is entirely consistent with the observational errors ($w = 0$) and all LB D abundances are consistent, within the errors, with \\ydlb~= $1.5(1 \\pm 0.03)$. This suggests that for the Local Bubble, \\ydism~$\\geq$ \\ydlb~and, $f_{\\rm D,LB} \\leq 1.8 \\pm 0.1$, consistent with all the successful chemical evolution models identified in SRT. However, while the uniform LB D abundance suggests that D may be {\\it undepleted} in the LB, for all LOS, \\ydmax~$\\approx 1.3$ \\ydlb, suggesting either that D is {\\it depleted uniformly} in the LB or, that outside of the LB the gas phase deuterium abundance may have been {\\it enhanced} along some LOS by the addition of nearly primordial gas which has recently fallen into the disk of the Galaxy in the form of cloudlets which take some time to mix with the pre-existing gas in the ISM. Does \\ydlb~= 1.5 or \\ydmax~= 2.0 provide the best estimate of the lower bound to the ISM D abundance? If deuterium is depleted onto dust, why is there not a strong correlation between deuterium abundance and iron depletion\\footnote{As pointed out by the Referee, shock strength may play a role in accounting for the scatter observed in the correlation between the gas phase D and Fe abundances. If deuterium is loosely bound to the grain mantle while iron is locked into the core of the dust grain, deuterium would be more easily returned to the gas than iron when grains are exposed to shocks of modest strength, while iron might be removed from dust grains only by stronger shocks. The scatter in the correlation between the gas phase D and Fe abundances may be an indicator of shock strength.} \\citep{linsky} and, which refractory element is then best to use as proxy for determining deuterium depletion onto dust? These questions cannot be answered by the analysis presented here. In a companion paper \\citep{sp}, abundances of refractory elements are used in concert with the deuterium abundances in an attempt to resolve this question. \\subsection*" }, "0910/0910.3770_arXiv.txt": { "abstract": "{ Baryonic Acoustic Oscillations (BAO) and their effects on the matter power spectrum can be studied by using the Lyman-$\\alpha$ absorption signature of the matter density field along quasar (QSO) lines of sight. A measurement sufficiently accurate to provide useful cosmological constraints requires the observation of $\\sim10^5$ quasars in the redshift range $2.2$90\\%) for QSOs with 0.3~$<$~$z$~$<$~2.2 but this completeness drops at higher redshift. The selection purity was brought up to 97\\% for $g<21$ using Kernel Density Estimation techniques applied to SDSS colors \\citep{kde04} and extended to the infrared by \\citet{richards09} implying that spectroscopy is not needed to confirm the corresponding statistical sample of quasars at high galactic latitudes. This led to the definition of a one-million-QSO catalog \\citep{kde09} down to $i=21.3$ from the photometry of SDSS Data Release 6 \\citep{adelman08}. Extending quasar selection methods to higher redshifts and magnitudes presents several difficulties. For example, at fainter magnitudes, galaxies start to contaminate ``point-like'' photometric catalogs both because of increasing photometric errors and because of non-negligible contributions of AGN's in certain bands. Nevertheless, such an extension is very desirable, not only to study the AGN population but also to use the quasars to study the foreground absorbers. In particular studies of spatial correlations in the IGM from the Lyman-$\\alpha$ forest and/or metal absorption lines are in need of higher target density at high redshift \\citep{petitjean97,nusser99,pichon01,caucci08}. More recently, it was realized that the Baryonic Acoustic Oscillations (BAO) could be detected in the Lyman-$\\alpha$ forest. BAO in the pre-recombination Universe imprint features in the matter power spectrum that have led to important constraints on the cosmological parameters. So far, BAO effects have been seen using galaxies of redshift $z<0.4$ to sample the matter density \\citep{eisenstein,cole,percival}. The Baryon Oscillation Spectroscopic Survey (BOSS) \\citep{boss} of the Sloan Digital Sky Survey (SDSS-III) \\citep{sdss3} proposes to extend these studies using galaxies of higher redshifts, $z<0.9$. The BOSS project will also study BAO effects in the range $2.22.2$ fainter QSOs ($\\sim 20$ QSOs per sq degree) and therefore requires the selection to be pushed up to $g\\sim 22$. We developed a new method to select quasars using more information than the standard color selection methods. The classification of objects is a task that is generally performed by applying cuts on various distributions which distinguish signal objects from background objects. This approach is not optimal because all the information (the shapes of the variable distributions, the correlations between the variables) is not exploited and this leads to a loss in classification efficiency. Statistical methods based on multivariate analysis have been developed to tackle this kind of problem. For historical reasons these methods have been focused on linear problems which are easily tractable. In order to deal with nonlinearities, Artificial Neural Networks (NN) have been shown to be a powerful tool in the classification task (see for instance \\citet{bishop}). By combining photometric measurements such as the magnitude values and their errors for the five bands ($ugriz$) of SDSS photometry, a NN approach will allow us both to select the QSO candidates and to predict their redshift. Similar methods such as Kernel Density Estimation (KDE) \\citep{kde04,kde09} already exist to select QSOs. Our approach based on NN is an extension of these methods because we will use more information (errors and absolute magnitude $g$ instead of only colors (difference between two magnitudes)). Moreover, we propose to treat in parallel the determination of the redshift with the same tool. This approach contrasts with the usual methods to compute photometric redshift which deal with $\\chi^2$ minimization techniques \\citep{photoz,weinstein04}. \\section {QSO and Background Samples} The quasar candidates should be selected among a photometric catalog of objects including real quasars and what we will call background objects. Here, both for the background and QSO samples, the photometric information comes from the SDSS-DR7 imaging database of point-like objects \\citep{dr7}, PLOs. We apply the same quality cuts on the photometry for the two samples and select objects with $g$ magnitude in the range $18 \\leq g \\leq 22$. Note that in the following, magnitudes will be point spread function (PSF) magnitudes \\citep{lupton99} in the SDSS pseudo-AB magnitude system \\citep{oke83}. \\subsection{Background Sample} \\label{sec:qsosample} For the background sample, we would ideally use an unbiased sample of spectroscopically confirmed SDSS point-like objects \\emph{that are not QSOs}. Unfortunately, we have no unbiased sample of such objects because spectroscopic targets were chosen in SDSS-I to favor particular types of objects. Fortunately, the number of QSOs among PLOs is sufficiently small that using all PLOs as background does not affect the NN's ability to identify QSOs. We have verified that this strategy works by using the synthetic PLO catalog of \\citet{fan}. We degraded the star sample by adding a few percent of QSOs in it. then, we retrained the NN and we compared the NN trained with a pure star sample. We did not observe any significant worsening of the NN performances. The background sample used in the following was drawn from the SDSS PLO sample. We used objects with galactic latitude $b$ around $45^\\circ$ to average the effect of Galactic extinction. In the future, we may consider the possibility of having a different NN for each stripe of constant galactic latitude. The final sample contains ~30,000 PLOs: half of them constituting the ``training\" sample, the other half the ``control\" sample, as explained in the next section. \\subsection{QSO Sample} For the QSO training sample, we use a list of 122,818 spectroscopically-confirmed quasars obtained from the 2QZ quasar catalog \\citep{croom04}, the SDSS-2dF LRG and QSO Survey (2SLAQ) \\citep{croom09}, and the SDSS-DR7 spectroscopic database \\citep{dr7}. These quasars have redshifts in the range $ 0.05 \\leq z \\leq 5.0 $ and $g$ magnitudes in the range $18 \\leq g \\leq 22$ (galactic extinction corrected). Since quasars will be observed over a limited blue wavelength range (down to about 3700~\\AA), we will target only quasars with $z>2.2$. Therefore, the sample of known quasars includes 33,918 QSOs with $z\\geq 1.8$: half of them constituting the effective ``training\" sample, the other half the ``control\" sample. For the determination of the photometric redshift, we use a wider sample of 95,266 QSOs with $z\\geq 1$. In order to compare together QSOs with background objects from different regions of the sky, the QSO magnitudes have been corrected for Galactic extinction with the model of \\citet{schlegel98}. \\subsection{Discriminating variables} The photometric information is extracted from the SDSS-DR7 imaging database \\citep{dr7}. The 10 elementary variables are the PSF magnitudes for the 5 SDSS bands ($ugriz$) and their errors. As explained in \\citet{kde09}, the most powerful variables are the four usual colors ($u-g$,$g-r$,$r-i$,$i-z$) which combine the PSF magnitudes. Fig.~\\ref{fig:Colors} shows the 2D color-color distributions for the QSO and PLO samples. These plots give the impression that it is easy to disentangle the two classes of objects but one needs to keep in mind that the final goal is to obtain a 50\\% efficiency for QSOs with a non-quasar PLO efficiency of the order of $\\sim10^{-3}$. Therefore to improve the NN performances, we added the absolute magnitude $g$ and the five magnitude errors. Their distributions for the two classes are given on Fig.~\\ref{fig:compVar}. An improvement can be expected from the additional variables and also from the correlations between the variables. Indeed, for example, it is expected that errors be larger for compact galaxies compared to intrinsic point-like objects. Note that the $g$ distribution for the QSOs is likely to be biased by the spectroscopic selection. This issue will be addressed in the future with the first observations of BOSS. Indeed the photometric selection of QSOs for these first observations is based on loose selection criteria and it should provide a ``less biased\" catalog of spectroscopically confirmed quasars, close to completeness up to $g=22$. ", "conclusions": "In this paper we have presented a new promising approach to select quasars from photometric catalogs and to estimate their redshift. We use an Neurone Network with a multilayer perceptron architecture. The input variables are photometric measurements, i.e. the magnitudes and their errors for the five bands ($ugriz$) of the SDSS photometry. For the target selection, we achieve a PLO rejection factor of 99.6\\% and 98.5\\% for, respectively, a quasar efficiency of 50\\% and 85\\%. The rms of the difference between the photometric redshift and the spectroscopic redshift is of the order of 0.15 over the region relevant for BAO studies. These new statistical methods developed in the context of the BOSS project can easily be extended to any other analysis requiring QSO selection and/or determination of their photometric redshift." }, "0910/0910.1557_arXiv.txt": { "abstract": "{} {We establish the mean metallicity from high-resolution spectroscopy for the recently found dwarf spheroidal galaxy Bo\\\"otes\\,I and test whether it is a common feature for ultra-faint dwarf spheroidal galaxies to show signs of inhomogeneous chemical evolution (e.g. as found in the Hercules dwarf spheroidal galaxy).} {We analyse high-resolution, moderate signal-to-noise spectra for seven red giant stars in the Bo\\\"otes\\,I dSph galaxy using standard abundance analysis techniques. In particular, we assume local thermodynamic equilibrium and employ spherical model atmospheres and codes that take the sphericity of the star into account when calculating the elemental abundances.} {We confirm previous determinations of the mean metallicity of the Bo\\\"otes\\,I dwarf spheroidal galaxy to be --2.3 dex. Whilst five stars are clustered around this metallicity, one is significantly more metal-poor, at --2.9 dex, and one is more metal-rich at, --1.9 dex. Additionally, we find that one of the stars, Boo-127, shows an atypically high [Mg/Ca] ratio, indicative of stochastic enrichment processes within the dSph galaxy. Similar results have previously only been found in the Hercules and Draco dSph galaxies and appear, so far, to be unique to this type of galaxy. } {} ", "introduction": "Until recently, the number of dwarf spheroidal (dSph) galaxies around the Milky Way was small compared to expectations from $\\Lambda$CDM \\citep{moore1999}. However, in the past few years several new systems have been found through systematic searches \\citep[e.g.][]{belokurov.boo,belokurovcats}. In general, dSph galaxies are some of the most tenuous stellar systems that we know of. This is especially true for the newly found dSph galaxies which have very low stellar luminosities \\citep[see e.g., ][]{martin08}. The new dSphs are ultra-faint and show low metallicities as indicated by low-resolution spectroscopy \\citep{kirbyletter,kochbierman}. \\citet{koch2008Her} found unusual abundance patterns in two red giant stars (RGB) in the ultra-faint Hercules dSph galaxy. Because of the low baryonic mass for these system it has been speculated that the elemental abundances in the stars in these systems might show us the results of individual supernova events \\citep{koch2008Her}. The recently found dSph galaxy in Bo\\\"otes \\citep[Bo\\\"otes\\,I, ][]{belokurov.boo} provides an excellent opportunity to test whether or not unusual elemental abundance ratios are a common feature of ultra-faint dSph galaxies, thanks to its low baryonic mass, \\citet{belokurov.boo} estimate, based on a colour magnitude diagram, that the Bo\\\"otes\\,I dSph galaxy is a purely old and metal-poor system. Low-resolution spectroscopic data confirm this \\citep{martin07,norris2008} at find $<$[Fe/H]$>$=--2.5. With $M_{\\rm V} \\sim -5.8$ this dSph galaxy is one of the least luminous galaxies known \\citep{belokurov.boo}. \\citet{fellhauer2008}, modelled the system and find that if this galaxy ever had a dark matter halo, it must still have it. This implies that, since the dark matter provides a deep potential well, the stars that originally formed in the dSph galaxy are still there and, moreover, the depth of the well should have helped retain the ejecta from core-collapse supernova. For Hercules, \\citet{koch2008Her} conclude that about 10 supernova are needed to pollute the interstellar medium to the observed atypical abundance ratios. Given that Bo\\\"otes\\,I has an even lower baryonic mass than Hercules, we might expect to be able to see enrichment from individual supernovae in the elemental abundance trends (which would show up as large scatter in element ratios from star to star). We have obtained high-resolution, moderate S/N spectra for seven RGB stars in the Bo\\\"otes\\,I dSph galaxy. Here we report on the mean metallicity, the metallicity spread, and atypical abundance ratios similar to those found in the Hercules \\citep{koch2008Her} and Draco dSph galaxies \\citep{fulbright2004Draco}. Thus, Bo\\\"otes\\,I becomes the third system to show unexpected abundances ratios. ", "conclusions": "" }, "0910/0910.4975_arXiv.txt": { "abstract": "We report on the method developed by Zibetti, Charlot \\& Rix (2009) to construct resolved stellar mass maps of galaxies from optical and NIR imaging. Accurate pixel-by-pixel colour information (specifically $g-i$ and $i-H$) is converted into stellar mass-to-light ratios with typical accuracy of 30\\%, based on median likelihoods derived from a Monte Carlo library of 50,000 stellar population synthesis models that include dust and updated TP-AGB phase prescriptions. Hence, surface mass densities are computed. In a pilot study, we analyze 9 galaxies spanning a broad range of morphologies. Among the main results, we find that: i) galaxies appear much smoother in stellar mass maps than at any optical or NIR wavelength; ii) total stellar mass estimates based on unresolved photometry are biased low with respect to the integral of resolved stellar mass maps, by up to 40\\%, due to dust obscured regions being under-represented in global colours; iii) within a galaxy, {\\it on local scales} colours correlate with surface stellar mass density; iv) the slope and tightness of this correlation reflect/depend on the morphology of the galaxy. ", "introduction": "The key role of stellar mass to determine (or predict) the physical properties of present day galaxies has been established by a number of works since the last decade \\citep[e.g.][]{gavazzi+96,scodeggio+02,kauffmann+03b}. Indication has been provided that stellar mass density might be even more fundamental \\citep{bell_dejong00,kauffmann+03b}. All these works, however, deal with {\\it global} estimates of the stellar mass-to-light ratio, which is assumed to be uniform throughout a galaxy, contrary to what we should expect based on well known spatial variations of stellar population and dust properties. Hence, resolving the distribution of stellar mass is crucial to properly measure total stellar mass and mass density and to start investigating the structure and dynamics of galaxies in an unbiased way. Having access to the mass distribution also allows to address questions like: Is there any relation between {\\it local} stellar mass density and the {\\it local} SED and physical properties {\\it within} a galaxy, similarly to the relations observed {\\it globally}? If so, what does it tell us about the internal mechanisms of galaxy evolution? With these goals and questions in mind, we have developed a new method to build stellar mass maps of galaxies \\citep[][ZCR09 hereafter]{ZCR09}, which we review in this contribution (Sec. 2). We also present preliminary results on the colour-stellar mass density relation within galaxies for a small sample of different morphological types (Sec. 3). ", "conclusions": "" }, "0910/0910.1885_arXiv.txt": { "abstract": "The five dimensional Brans-Dicke theory naturally provides two scalar fields by the Killing reduction mechanism. These two scalar fields could account for the accelerated expansion of the universe. We test this model and constrain its parameter by using the type Ia supernova (SN Ia) data. We find that the best fit value of the 5-dimensional Brans-Dicke coupling contant is $\\omega = -1.9$. This result is also consistent with other observations such as the baryon acoustic oscillation (BAO). ", "introduction": "The expansion of the Universe is shown to be accelerating by the observations of type Ia supernovae (SN Ia) \\cite{Perlmutter99, Riess98}. This is usually attributed to the contribution of an unknown component, dubbed dark energy, which has negative pressure and makes up about three quarters of the total cosmic density (for recent measurements, see e.g. Ref.~\\cite{Komatsu08,Gong08}). The simplest model for dark energy is the cosmological constant (CC), which is consistent with most of the observations today. However, there are two big problems for CC, i.e. the well-known ``fine tuning problem'' and the ``coincidence problem'' . As alternatives, and also to solve these two problems, many dynamical dark energy models with scalar field have been proposed, such as quintessence \\cite{Wetterich88,Ratra88,FJ97,CDS98,CLW98, Carroll98,ZMS99,ZMS991}, phantom \\cite{Caldwell02}, quintom\\cite{FWZ05}, K-essence \\cite{Chiba00, Armendariz00}, tachyon \\cite{Padmanabhan02, Bagla03,Abramo03,Agui04,Guo04,Copeland05} and so on. Nevertheless, in most cases the fundamental physical origin of these scalar fields remain unknown, but just added by hand. In Ref. \\cite{Qiang05}, by generalizing the Brans-Dicke theory to five dimensions and exploring its effect on the 4-dimensional world, another interesting approach to explain the cosmic accelerated expansion was proposed. Under the condition that the extra dimension is compact and sufficiently small, a spacelike Killing vector field $\\xi^a$ arises naturally, in which case the 5D Brans-Dicke theory can be reduced to a 4D theory, such that the 4-metric is coupled with two scalar fields $\\phi$ and $\\lambda$. Notes that here the scalar fields in four dimension stem naturally from a fundamental theory of gravity. Considering the hypersurface-orthogonal property of $\\xi^a$, the line element in five dimension can take the form as \\begin{equation} ds^2=g_{\\mu\\nu}dx^{\\mu}dx^{\\nu}+\\lambda dx^5dx^5,\\label{metric} \\end{equation} thus the scalar $\\lambda$ also plays the role of a ``scale factor'' of the extra dimension. It was shown in Ref. \\cite{Qiang05} that these two scalar fields originated from the Killing reduction of the 5D Brans-Dicke theory may lead to the accelerated expansion of the universe. More detailed analysis is desirable to check if the theory can match the current observational data such as the SN Ia and the baryon acoustic oscillation. In this paper, we compare the predictions of the cosmic expansion rate of this theory with the current cosmological observations, and constrain the 5D Brans-Dicke theory by means of the SN Ia data and the baryon acoustic oscillation (BAO) measurements. This work is based on the fact that the 4-dimensional gravitational constant $G$ varies extremely slowly with time \\cite{Turyshev04} in the current epoch. Thus we assume that $G$ is a constant at the ``low redshift'', so that the accretion of the white dwarf will not be affected, hence the luminosity and light curve of the observed SN Ia (redshifts range from 0 to 2) are not affected by the slow variations of $G$, and the SN Ia can still be used as the standard candle. Furthermore, if we assume that $G$ is almost a constant throughout the history of the Universe, we could also use the information from the large scale structure (e.g. BAO) to perform the constraints. However, we should note that $G=(\\phi\\lambda^{1/2}L)^{-1}$ \\cite{Qiang05}, where $L$ is the coordinate scale of the extra dimension. Although $G$ is almost constant in four dimension, $G^{(5)}\\sim\\phi^{-1}$ is not necessarily a constant and can still evolve with time. ", "conclusions": "By considering a hypersurface-orthogonal spacelike Killing vector field in the 5-dimensional spacetime, the 5D Brans-Dicke theory can be reduced to a 4D theory with the 4-metric coupled to two scalar fields. These two fields could naturally lead to the accelerated expansion of the Universe. We study the evolution of the two fields and compare the expansion rate with SN Ia observations. The two scalar field would make the Universe evolve as if ``matter-dominate'' or ``dark energy-dominate'' when $\\omega$ is greater or less than -2. We find that the model is in best agreement with the supernovae data when the 5-dimensional coupling constant $\\omega =-1.9 \\approx -2$, which happens to be also the value required to satisfy the solar system experiments. Furthermore, for this best fit value, the best fit $\\Omega_{m_0}$ value is about 0.27, in good agreement with other independent measurements such as those derived from X-ray cluster observations. This work is based on the assumption that the 4D gravitational constant $G$ varies extremely slowly so that it can be regarded as a constant at \"low redshift\" where the SN Ia data are available. If we further assume $G$ does not change during the whole history of the Universe, then other cosmological observations such as BAO can also be used, we find that in this case the results are almost the same. In conclusion, the 5-dimensional Brans-Dicke theory could naturally provide two scalar fields which may cause the accelerated expansion, the result is consistent with the SN Ia observation, hence it is a candidate to explain the accelerated expansion of the Universe." }, "0910/0910.3552_arXiv.txt": { "abstract": "A solution to the ultra-relativistic strong explosion problem with a non-power law density gradient is delineated. We consider a blast wave expanding into a density profile falling off as a steep radial power-law with small, spherically symmetric, and log-periodic density perturbations. We find discretely self-similar solutions to the perturbation equations and compare them to numerical simulations. These results are then generalized to encompass small spherically symmetric perturbations with arbitrary profiles. ", "introduction": "The profusion of explosions occurring in the observable universe has led to a pointed interest in the dynamics of blast waves. The quantitative treatment of strong explosions began with the self-similar solutions found by Sedov, Von-Neumann and Taylor \\cite{Sedov} \\cite{VonNeumann} \\cite{Taylor} for the flow behind spherical Newtonian shocks propagating into a cold gas with a power-law density profile, $\\rho \\propto r^{-k}$. This solution is valid for moderately steep decay exponents, $k<3$. Subsequently, corresponding solutions were found in the ultra-relativistic regime by Blandford \\& McKee \\cite{BM} for $k<4$. In these solutions the Lorentz factor of the shock $\\Gamma$ scales as $\\Gamma^2 \\propto t^{-m}$, and $m$ is fixed by energy conservation arguments. For $k>4$ this procedure fails due to the energy in the Blandford-McKee solution diverging, and a different argument must be used if a self-similar solution is to be found. It turns out that there exists a $k_g>4$ such that for $k>k_g$ just such an argument exists \\cite{Best}. The solutions in this regime are called type-II solutions \\cite{Zeldovich}, and what sets them apart from the solutions with $k<4$, known as type-I solutions, is the rapid acceleration of the shock and the fluid behind it that causes the formation of a sonic point between the shock and the center of the explosion. This point (actually a spherical surface) marks the boundary of an inner region that becomes causally disconnected from the shock. This allows a self-similar flow behind the shock to coexist with a non self-similar flow further away from it, thus maintaining a finite amount of total energy. For the ultra-relativistic case, $k_g=5-\\sqrt{3/4} \\approx 4.13$. The sonic point appears as a singularity in the hydrodynamic equations, and the requirement of regularity in traversing this singularity supplies the necessary condition to fix the value of $m$. In this paper we focus on Type-II ultra-relativistic solutions with $k>k_g$ \\cite{Best}, and use these as the basis for a perturbative analysis where we disturb the external density profile. The basic method for doing so was developed in a previous paper \\cite{Oren} for the Newtonian case, and we adapt it here for the ultra-relativistic case. We introduce a special family of perturbations to the ambient medium surrounding the explosion, so that we are able to reduce the hydrodynamic equations to a set of ordinary differential equations, and then find the flow behind the shock in the presence of perturbations. In section \\ref{sec:unpert} we briefly describe the unperturbed solutions, while in section \\ref{sec:RI} we shed light on some general properties of these solutions. In section \\ref{sec:pert} we write down the equations for the perturbations and solve them, and in section \\ref{sec:fourier} we generalize the results of the previous section by using a spectral decomposition of an arbitrary perturbation profile. Finally in section \\ref{sec:discuss} we summarize and discuss our results. ", "conclusions": "\\label{sec:discuss} We have laid out a method of solving the hydrodynamic equations for a strong explosion in the presence of spherically symmetric perturbations to the ambient density. At first we study a special group of perturbations with log-sinusoidal radial dependence, and discover an analytic solution. The requirement of spherical symmetry is not easily relaxed, because relativistic effects make it difficult to find self-similar solutions with a non trivial angular dependence. Another limitation is the linearity of the perturbation analysis that limits the validity of the solutions to small amplitudes. The smallness required may be seen in figure \\ref{fig:sigmas} where numerical solutions with different amplitudes are superimposed against the analytical solution. It can be seen that our theory gives reasonably accurate results for amplitudes up to about $0.1$. The nonlinearity of waves with higher amplitudes is expressed through shock formation before their crests, as can be seen in the red line in figure \\ref{fig:sigmas}. On the other hand we take advantage of linearity to generalize our results using a Fourier-like method, decomposing an arbitrary perturbation to simple modes for which we can solve the equations analytically. In this way we can treat interesting scenarios like a sudden rise or drop in density, which might e.g. be encountered in a stellar wind due to interaction with the interstellar medium. Another possible application is the emergence of a shock from the edge of a star \\cite{Pan}, where the drop in density accelerates shocks to relativistic velocities. The effect of perturbations in the star's envelope can be treated with the method presented here, requiring only an adaptation of the unperturbed solution. Acknowledgments: The authors wish to thank Prof. Tsvi Piran for fruitful discussions. This research was partially supported by a NASA grant, IRG grant and a Packard Fellowship." }, "0910/0910.2138_arXiv.txt": { "abstract": "\\label{abstract}\\label{firstpage} The nature of the progenitors of Type Ia supernovae (SNe Ia) is still unclear. In this paper, by considering the effect of the instability of accretion disk on the evolution of white dwarf (WD) binaries, we performed binary evolution calculations for about 2400 close WD binaries, in which a carbon--oxygen WD accretes material from a main-sequence star or a slightly evolved subgiant star (WD + MS channel), or a red-giant star (WD + RG channel) to increase its mass to the Chandrasekhar (Ch) mass limit. According to these calculations, we mapped out the initial parameters for SNe Ia in the orbital period--secondary mass ($\\log P^{\\rm i}-M^{\\rm i}_2$) plane for various WD masses for these two channels, respectively. We confirm that WDs in the WD + MS channel with a mass as low as $0.61\\,M_\\odot$ can accrete efficiently and reach the Ch limit, while the lowest WD mass for the WD + RG channel is $1.0\\,\\rm M_\\odot$. We have implemented these results in a binary population synthesis study to obtain the SN Ia birthrates and the evolution of SN Ia birthrates with time for both a constant star formation rate and a single starburst. We find that the Galactic SN Ia birthrate from the WD + MS channel is $\\sim$$1.8\\times 10^{-3}\\ {\\rm yr}^{-1}$ according to our standard model, which is higher than previous results. However, similar to previous studies, the birthrate from the WD + RG channel is still low ($\\sim$$3\\times 10^{-5}\\ {\\rm yr}^{-1}$). We also find that about one third of SNe Ia from the WD + MS channel and all SNe Ia from the WD + RG channel can contribute to the old populations ($\\ga$1\\,Gyr) of SN Ia progenitors. ", "introduction": "Type Ia supernovae (SNe Ia) play an important role in astrophysics, especially in cosmology. They appear to be good cosmological distance indicators and have been applied successfully to the task of determining cosmological parameters (e.g. $\\Omega$ and $\\Lambda$: Riess et al. 1998; Perlmutter et al. 1999). They are also a key part of our understanding of galactic chemical evolution owing to the main contribution of iron to their host galaxies (e.g. Greggio \\& Renzini 1983; Matteucci \\& Greggio 1986). Despite their importance, several key issues related to the nature of their progenitors and the physics of the explosion mechanisms are still not well understood (Hillebrandt \\& Niemeyer 2000; R\\\"{o}pke \\& Hillebrandt 2005; Podsiadlowski et al. 2008; Wang et al. 2008a), and no SN Ia progenitor system has been conclusively identified pre-explosion. There is a broad agreement that SNe Ia are thermonuclear explosions of carbon--oxygen white dwarfs (CO WDs) in binaries (see the review of Nomoto, Iwamoto \\& Kishimoto 1997). Over the last few decades, two competing progenitor models of SNe Ia were discussed frequently, i.e. the double-degenerate (DD) and single-degenerate (SD) models. Of these two models, the SD model (Whelan \\& Iben 1973; Nomoto, Thielemann \\& Yokoi 1984; Fedorova, Tutukov \\& Yungelson 2004; Han 2008) is widely accepted at present. It is suggested that the DD model, which involves the merger of two CO WDs (Iben \\& Tutukov 1984; Webbink 1984; Han 1998), likely leads to an accretion-induced collapse rather than a SN Ia (Nomoto \\& Iben 1985; Saio \\& Nomoto 1985; Timmes, Woosley \\& Taam 1994). For the SD model, the companion is probably a main-sequence (MS) star or a slightly evolved subgiant star (WD + MS channel), or a red-giant star (WD + RG channel), or even a He star (WD + He star channel) (Hachisu, Kato \\& Nomoto 1996; Li $\\&$ van den Heuvel 1997; Hachisu et al. 1999a; Langer et al. 2000; Han $\\&$ Podsiadlowski 2004, 2006; Wang et al. 2009a,b). Although the SD model is currently a favorable progenitor model of SNe Ia, any single channel of the model cannot account for the birthrate inferred observationally. Note that, some recent observations have indirectly suggested that at least some SNe Ia can be produced by a variety of different progenitor systems (e.g. Hansen 2003; Ruiz-Lapuente et al. 2004; Patat et al. 2007; Voss \\& Nelemans 2008; Wang et al. 2008b; Justham et al. 2009). At present, various progenitor models of SNe Ia can be examined by comparing the distribution of the delay time (between the star formation and SN Ia explosion) expected from a progenitor channel with that of observations (e.g. Chen $\\&$ Li 2007; Meng, Chen \\& Han 2009; L\\\"{u} et al. 2009; Ruiter, Belczynski \\& Fryer 2009). Mannucci, Della Valle \\& Panagia (2006) argued for the existence of two separate SN Ia populations, a `prompt' component with a delay time less than $\\sim$100\\,Myr, and a `delayed' component with a delay time $\\sim$3\\,Gyr. By investigating the star formation history (SFH) of 257 SN Ia host galaxies, Aubourg et al. (2008) found evidence of a short-lived population of SN Ia progenitors with lifetimes less than 180\\,Myr. Botticella et al. (2008) and Totani et al. (2008) analyzed host galaxies of SNe Ia and concluded that a substantial fraction of SNe Ia must have long delay times on the order of 2$-$3\\,Gyr. Moreover, Schawinski (2009) recently constrained the minimum delay time of 21 nearby SNe Ia by investigating their host galaxies (early-type galaxies). The study showed that these early-type host galaxies lack `prompt' SNe Ia with a delay time less than $\\sim$100\\,Myr and that $\\sim$70 per cent SNe Ia have minimum delay times of 275\\,Myr$-$1.25\\,Gyr, while at least 20 per cent SNe Ia have minimum delay times of at least 1\\,Gyr at 95 per cent confidence and two of these four SNe Ia are likely older than 2\\,Gyr. For SNe Ia with the short delay times, Wang et al. (2009a) studied a WD + He star channel to produce SNe Ia, in which a CO WD accretes material from a He MS star or a He subgiant to increase its mass to the Chandrasekhar (Ch) mass. The study showed the parameter spaces for the progenitors of SNe Ia. By using a detailed binary population synthesis (BPS) approach, Wang et al. (2009b, WCMH09) found that the Galactic SN Ia birthrate from this channel is $\\sim$$0.3\\times 10^{-3}\\ {\\rm yr}^{-1}$, and that this channel can produce SNe Ia with short delay times ($\\sim$45$-$140\\,Myr). For SNe Ia with the long delay times ($\\ga$1\\,Gyr), this requires that the mass of the companion should be $\\la$$2\\,M_{\\odot}$. Recently, Xu \\& Li (2009) emphasized that the mass-transfer through the Roche lobe overflow (RLOF) in the evolution of WD binaries may become unstable (at least during part of the mass-transfer lifetime), i.e. the mass-transfer rate is not equivalent to the mass-accretion rate onto the WD. This important feature has been ignored in nearly all of the previous theoretical works on SN Ia progenitors except for King, Rolfe \\& Schenker (2003) \\footnote{King, Rolfe \\& Schenker (2003) adopted a similar method in Li \\& Wang (1998) to produce SNe Ia with long period dwarf novae in a semi-analytic approach.} and Xu \\& Li (2009), who inferred that the mass-accretion rate onto the WD during dwarf nova outbursts can be sufficiently high to allow steady nuclear burning of the accreted matter and growth of the WD mass. In particular, the study of Xu \\& Li (2009) enlarges the region of SN Ia parameter spaces. However, they only give the SN Ia parameter spaces with WD initial masses of 0.8 and $1\\,M_{\\odot}$. More detailed work is obviously needed to investigate the influence of the accretion disk on the final results and to give SN Ia birthrates and delay times. Including the effect of the instability of accretion disk on the evolution of WD binaries, the purpose of this paper is to study the WD + MS and WD + RG channels towards SNe Ia comprehensively and systematically, and then to determine the parameter spaces for SNe Ia, which can be used in BPS studies. In Section 2, we describe the numerical code for the binary evolution calculations and the grid of the binary models. The binary evolutionary results are shown in Section 3. We describe the BPS method in Section 4 and present the BPS results in section 5. Finally, a discussion is given in Section 6, and a summary in Section 7. ", "conclusions": "\\label{6. DISCUSSION} The regions (Fig. 6 in this paper) for producing SNe Ia depend on many uncertain input parameters, in particular for the duty cycle which is poorly known. The main uncertainties lie in the facts that it varies from one binary system to another and may evolve with the orbital periods and mass-transfer rates (e.g. Lasota 2001; Xu \\& Li 2009). This is the reason why we choose an intermediate value (0.01) rather than other extreme ones (e.g. 0.1 or $10^{-3}$). However, we also did some tests for a higher or lower value of the duty cycle. We find that the variation of the duty cycle value will influence the regions for producing SNe Ia, especially for the WD + RG channel (for a high value (0.1), the regions will be shifted to higher period; for a low value ($10^{-3}$), the regions will be shifted to lower period). For these two extreme duty cycle values, the SN Ia birthrate from the WD + RG channel will be lower due to the smaller regions for producing SNe Ia. In our binary evolution calculations we have not considered the influence of rotation on the H-accreting WDs. The calculations of Yoon, Langer \\& Scheithauer (2004) showed that the He-shell burning is much less violent when rotation is taken into account. This may significantly increase the He-accretion efficiency ($\\eta _{\\rm He}$ in this paper). Meanwhile, the maximum stable mass of a rotating WD may be above the standard Ch mass (i.e. the super-Ch mass model: Uenishi, Nomoto \\& Hachisu 2003; Yoon \\& Langer 2005; Chen \\& Li 2009). However, we mainly focus on the standard Ch mass explosions of the accreting WDs in this work. In our BPS studies we assume that all stars are in binaries and about 50\\,per cent of stellar systems have orbital periods less than 100\\,yr. In fact, this is known to be a simplification, and the binary fractions may depend on metallicity, environment, spectral type, etc. If we adopt 40\\,per cent of stellar systems have orbital periods below 100\\,yr by adjusting the parameters in Equation (10), we estimate that the Galactic SN Ia birthrate from the WD + MS channel will decrease to be $\\sim$$1.4\\times 10^{-3}\\ {\\rm yr}^{-1}$ according to our standard model (the birthrate from the WD + RG channel will decrease to be $\\sim$$2\\times 10^{-5}\\ {\\rm yr}^{-1}$). In addition, Umeda et al. (1999) concluded that the upper limit mass of CO cores born in binaries is about 1.07$\\,M_\\odot$. If this value is adopted as the upper limit of the CO WD, the SN Ia birthrate from the WD + MS channel will decrease to be $\\sim$$1.7\\times10^{-3}\\,{\\rm yr}^{-1}$ (the birthrate from the WD + RG channel will decrease to be $\\sim$$1\\times 10^{-5}\\ {\\rm yr}^{-1}$). The Galactic SN Ia birthrate from the WD + RG channel is $\\sim$3$\\times 10^{-5}\\ {\\rm yr}^{-1}$ according to our standard model, which is low compared with observations, i.e. SNe Ia from this channel may be rare. However, further study on this channel is necessary, since this channel may explain some SNe Ia with long delay times. In addition, it is suggested that, RS Oph and T CrB, both recurrent novae are probable SN Ia progenitors and belong to the WD + RG channel (e.g. Belczy$\\acute{\\rm n}$ski \\& Mikolajewska 1998; Hachisu, Kato \\& Nomoto 1999b; Sokoloski et al. 2006; Hachisu, Kato \\& Luna 2007). Meanwhile, by detecting Na I absorption lines with low expansion velocities, Patat et al. (2007) suggested that the companion of the progenitor of SN 2006X may be an early RG star. Additionally, Voss \\& Nelemans (2008) studied the pre-explosion archival X-ray images at the position of the recent SN 2007on, and they considered that its progenitor may be a WD + RG system. Employing Eggleton's stellar evolution code with the prescription of Hachisu et al. (1999a) for the mass-accretion of CO WDs, and including the effect of the instability of accretion disk on the evolution of WD binaries, we performed binary evolution calculations for about 2400 close WD binaries. The calculated results further confirm that the disk instability could substantially increase the mass-accumulation efficiency for accreting WDs, and cause the possible SNe Ia to occur in systems with $\\la$$2\\,M_{\\odot}$ donor stars (see also King, Rolfe \\& Schenker 2003; Xu \\& Li 2009). We find that the Galactic SN Ia birthrate from the WD + MS channel is $\\sim$$1.8\\times 10^{-3}\\ {\\rm yr}^{-1}$ according to our standard model, which is higher than previous results. However, similar to previous studies, the birthrate from the WD + RG channel is still low ($\\sim$$3\\times 10^{-5}\\ {\\rm yr}^{-1}$). We also find that about one third of SNe Ia from the WD + MS channel and all SNe Ia from the WD + RG channel can contribute to the old populations ($\\ga$1\\,Gyr) of SN Ia progenitors. The companion stars of SNe Ia with long delay times in this work would survive in the SN explosion and show distinguishing properties. In future investigations, we will explore the properties of the companion stars after SN explosion, which could be verified by future observations." }, "0910/0910.5718_arXiv.txt": { "abstract": "We present a study of galaxy mergers and the influence of environment in the Abell 901/902 supercluster at $z\\sim$~0.165, based on 893 bright ($R_{\\rm Vega} \\le$~24) intermediate mass ($M_{*} \\geq 10^{9} M_{\\sun}$) galaxies. We use $HST$ ACS F606W data from the Space Telescope A901/902 Galaxy Evolution Survey (STAGES), COMBO-17, $Spitzer$ 24$\\micron$, and $XMM$-$Newton$ X-ray data. Our analysis utilizes both a physically driven visual classification system, and quantitative CAS parameters to identify systems which show evidence of a recent or ongoing merger of mass ratio $ >$~1/10 (i.e., major and minor mergers). Our results are: (1)~After visual classification and minimizing the contamination from false projection pairs, we find that the merger fraction $f_{\\rm merge}$ is 0.023$\\pm$0.007. The estimated fractions of likely major mergers, likely minor mergers, and ambiguous cases are 0.01$\\pm$0.004, 0.006$\\pm$0.003, and 0.007$\\pm$0.003, respectively. (2)~All the mergers lie outside the cluster core of radius $R <$~0.25 Mpc: the lack of mergers in the core is likely due to the large galaxy velocity dispersion in the core. The mergers, instead, populate the region (0.25 Mpc $< R\\leq$~2 Mpc) between the core and the cluster outskirt. In this region, the estimated frequency of mergers is similar to those seen at typical group overdensities in N-body simulations of accreting groups in the A901/902 clusters. This suggests ongoing growth of the clusters via accretion of group and field galaxies. (3)~We compare our observed merger fraction with those reported in other clusters and groups out to $z\\sim$~0.4. Existing data points on the merger fraction for $L \\leq L^{*}$ galaxies in clusters allow for a wide spectrum of scenarios, ranging from no evolution to evolution by a factor of $\\sim$5 over $z\\sim$~0.17 to 0.4. (4)~In A901/902, the fraction of mergers, which lie on the blue cloud is 80$\\%\\pm 18\\%$ (16/20) versus 34$\\%\\pm7\\%$ or (294/866) for non-interacting galaxies, implying that interacting galaxies are preferentially blue. (5)~The average SFR, based on UV or a combination of UV+IR data, is enhanced by a factor of $\\sim$1.5 to 2 in mergers compared to non-interacting galaxies. However, mergers in the A901/902 clusters contribute only a small fraction (between 10\\% and 15\\%) of the total SFR density, while the rest of the SFR density comes from non-interacting galaxies. ", "introduction": "Understanding how galaxies evolve in various environments (field, groups, and clusters), and as a function of redshift is a key step toward developing a coherent picture of galaxy evolution. Present-day cluster and field galaxies differ due to several factors, which are often grouped under the umbrella of `nature' versus `nurture'. First, in cold dark matter (CDM) cosmogonies, the first galaxies formed and evolved early in cluster cores, as the higher initial overdensities led to faster gravitational collapse and more rapid mergers of proto-galaxies (e.g., Cole et \\etal 2000; Steinmetz \\& Navarro 2002). Second, in the context of the bottom-up CDM assembly paradigm, the outer parts of clusters and superclusters, grow at late times via mergers, smooth accretion, and discrete accretion of groups and field galaxies. This idea is supported by observational studies (e.g., Zabludoff \\& Franx 1993; Abraham \\etal 1996a; Balogh, Navarro \\& Morris 2000), which suggest that clusters continuously grew by the accretion of groups. Third, the dominant physical processes affecting galaxies differ in cluster and field environments due to the different galaxy number density, galaxy velocity dispersion, and intracluster medium (ICM). Among these processes are close galaxy-galaxy interactions, such as strong tidal interactions and mergers (e.g., Barnes 1992; Moore \\etal 1998) and galaxy harassment (e.g., Moore \\etal 1996), which stems from the cumulative effect of weak interactions. Furthermore, in clusters where the hot ICM makes up as much as 15\\% of the total mass, galaxy-ICM interactions, such as ram pressure stripping (Gunn \\& Gott 1972; Larson \\etal 1980; Quilis \\etal 2000; Balogh, Navarro \\& Morris 2000), can play an important role in removing the diffuse gas from galaxies. The tidal field of the cluster potential may also play a relevant role in the dynamical evolution of cluster galaxies (Gnedin 2003). Systematic studies of the differences between cluster, group, and field galaxies at different redshifts are needed to shed light on the relative importance of these various processes in their respective environments. Several differences have been observed between galaxies in the field and those in the rich cluster environment, but the physical drivers behind these variations are still under investigation. At $z\\sim$~0, the relative percent of massive early type (E+S0) galaxies to spirals rises from (10\\%+10\\%:80\\%) in the field to (40\\%+50\\%:10\\%) in the cores of very rich clusters, leading to the so-called morphology density relation (Dressler 1980; Dressler \\etal 1997). However, recent Sloan Digital Sky Survey (SDSS) studies suggest that masses and star formation (SF) histories of galaxies are more closely related to environmental physical processes rather than their structural properties (Blanton \\etal 2005). The SF histories of galaxies depend on both luminosity (Cole et al. 2001) and environment (Diaferio \\etal 2001; Koopmann \\& Kenney 2004). The fraction of blue galaxies in clusters appears to rise with redshift, an effect known as the Butcher Oemler effect (Butcher \\& Oemler 1978; Margoniner \\etal 2001; de Propis \\etal 2003). There is also evidence that SF in bright ($M_{\\rm v} < $~-18) cluster galaxies is suppressed compared to field galaxies (e.g., Balogh \\etal 1998, 1999), for reasons that are not well understood. Galaxy interactions and mergers have been proposed as a mechanism for the change in galaxy populations in clusters from that of the field (e.g., Toomre \\& Toomre 1972; Lavery \\& Henry 1988; Lavery Pierce, \\& McClure 1993). There have been various studies on the properties of galaxies (e.g., Dressler 1980, Postman \\& Geller 1984, Giovanelli, Haynes, \\& Chincarini, 1986, Kennicutt 1983; Gavazzi \\& Jaffe 1985, Whitmore \\etal 1993) and of galaxy interactions and mergers (e.g., Lavery \\& Henry 1988; Lavery, Pierce, \\& Mclure 1992; Zepf 1993; Dressler \\etal 1994; Couch \\etal 1998; van Dokkum \\etal 1998, 1999; Tran \\etal 2005, 2008) in different environments. Some of these studies suggest that galaxy interactions and mergers may play a role in morphological transformations of galaxies in clusters, but there have been few systematic studies of galaxy mergers and interactions in clusters, based on high resolution $HST$ images as well as $Spitzer$ 24$\\micron$, and X-ray images. In this paper we present a study of the frequency, distribution, color, and SF properties of galaxy mergers in the A901/902 supercluster at $z\\sim$0.165. We use $HST$ ACS F606W data taken as part of the Space Telescope A901/902 Galaxy Evolution Survey (STAGES; Gray \\etal 2009), along with ground-based COMBO-17 imaging data (Wolf \\etal 2004), $Spitzer$ 24$\\micron$ data (Bell \\etal 2005, 2007), $XMM$-$Newton$ X-ray data (Gilmour \\etal 2007), and dark matter (DM) mass measurements from weak lensing (Heymans \\etal 2008). With a resolution of $0.1\\arcsec$ or $\\sim$280 pc at $z=0.165$, the $HST$ images allow for the identification of merger signatures such as double nuclei, arcs, shells, tails, tidal debris, and accreting satellites. The COMBO-17 survey (Wolf \\etal 2004) provides accurate spectrophotometric redshifts down to $R_{\\rm Vega}$ of 24 and stellar masses (Borch \\etal 2006). The $Spitzer$ 24$\\micron$ data (Bell \\etal 2005, 2007) probe the obscured SF, while X-ray maps (Gilmour \\etal 2007; Gray \\etal, 2010 in preparation) provide information of how the ICM density changes throughout the cluster. We present the data and sample selection in $\\S$Section~\\ref{sdatas}. In $\\S$Section~\\ref{smvc} and $\\S$Section~\\ref{smcas1}, we describe the two different methods that we use to identify galaxy mergers: a physically motivated classification system that uses visual morphologies, stellar masses, and spectrophotometric redshifts, and a method based on CAS asymmetry $A$ and clumpiness $S$ parameters (Conselice \\etal 2000). In $\\S$Section~\\ref{sinte1} and~\\ref{scas1}, we explore the frequency of galaxy mergers in A901/902 based on these two methods and present one of the first systematic comparisons to date between CAS-based and visual classification results in clusters. We set a lower limit on the fraction of major mergers (those with mass ratio $M_{1}/M_{2} \\ge 1/4$). In $\\S$Section~\\ref{sinte2}, we examine the distribution of mergers in the A901/902 supercluster as a function of clustocentric radius, galaxy number density, local galaxy surface density, relative ICM density, and local DM mass surface density. In $\\S$Section~\\ref{sinte3}, we compare our results on the fraction and distribution of mergers to expectations based on analytical estimates and simulations of mergers in different environments. In $\\S$Section~\\ref{scomp1}, we compare our results on galaxy mergers in the A901/902 supercluster to groups and clusters at different redshifts out to $z\\sim$~0.8. We investigate the fraction of mergers and non-interacting galaxies on the blue cloud and red sample as a function of clustocentric radius in $\\S$Section~\\ref{scolor1}. Finally in $\\S$Section~\\ref{ssfr1}, we compare the star formation rate (SFR) of mergers and non-interacting galaxies in the A901/902 clusters. The results of this work are summarized in $\\S$Section~\\ref{ssumm}. In this paper, we assume a flat cosmology with $\\Omega_{\\rm m}$ = 1 $-\\Omega_{\\lambda}$ = 0.3 and H$_{0}$~=~70 km s$^{-1}$ Mpc$^{-1}$ throughout this paper. ", "conclusions": "\\label{sresul} \\subsection{Merger fraction in A901/902 from visual classification}\\label{sinte1} Next we discuss how to define and estimate the merger fraction. Our goal is to estimate the fraction of $f_{\\rm merge}$ of systems with stellar mass above an appropriately chosen mass cut $M_{\\rm cut}$, which are involved in mergers of mass ratio $\\ge$~1/10. The merger fraction $f_{\\rm merge}$ is computed as ($N_{\\rm merge}/N_{\\rm tot}$), where $N_{\\rm merge}$ is the number of major and minor mergers involving galaxies with $M_{*} \\ge M_{\\rm cut}$, and $N_{\\rm tot}$ is the total number of galaxies with $M_{*} \\ge M_{\\rm cut}$. It is important to determine the minimum stellar masses of the major and minor mergers involving galaxies with $M_{*} \\ge M_{\\rm cut}$ and thereby assess for what value of $M_{\\rm cut}$ we can trace such mergers. Major merger pairs (defined as having mass ratio 1/4 $< M_{1}/M_{2} \\le$ 1) of mass ratio 1:1 to 1:3 will have minimum stellar masses ranging from $2 \\times M_{\\rm cut}$ to $4 \\times M_{\\rm cut}$, in cases where galaxies of mass $M_{*} \\ge M_{\\rm cut}$ merge with systems at least as massive as themselves. However, if galaxies of mass $M_{*} \\ge M_{\\rm cut}$ merge with lower mass systems, the minimum mass of 1:1 to 1:3 mergers will range from $2 \\times M_{\\rm cut}$ to ($4 \\times M_{\\rm cut}/3$). Taken together, the above constraints imply that we are complete for major mergers involving galaxies with $M_{*} \\ge M_{\\rm cut}$, as long as we can trace {\\rm 1:3 major mergers with a minimum stellar mass of ($4 \\times M_{\\rm cut}/3$). Repeating the exercise for minor mergers (defined as having 1/10 $< M_{1}/M_{2} \\le$ 1/4), it follows that the minimum stellar mass for 1:4 to 1:9 mergers ranges from $5 \\times M_{\\rm cut}$ to $9 \\times M_{\\rm cut}$ or from ($5 \\times M_{\\rm cut}/4$) to ($10 \\times M_{\\rm cut}/9$), depending on whether galaxies of mass $M_{*} \\ge M_{\\rm cut}$ merge with systems of higher mass or lower mass. Thus, we are complete for minor mergers involving galaxies with $M_{*} \\ge M_{\\rm cut}$, as long as we can trace {\\rm 1:9 minor mergers with a minimum stellar mass of ($10 \\times M_{\\rm cut}/9$). For what value of $M_{\\rm cut}$ are these conditions satisfied by the mergers of Types 1, 2a, and 2b, which we visually identified in the final sample ($\\S$Section~\\ref{smvc})? Since our final sample is complete for $M_{*} \\ge 10^{9} M_{\\sun}$, it follows that we can identify all mergers of type 1 and 2a with $M_{*} \\ge 10^{9} M_{\\sun}$ from this sample. If we impose this value to the afore-defined criteria of ($10 \\times M_{\\rm cut}/9$) and ($4 \\times M_{\\rm cut}/3$), it implies that a mass cut $ M_{\\rm cut} \\ge 0.9 \\times 10^{9} M_{\\sun}$ would allow us to be complete for major and minor mergers of type 1 and 2a. However, the situation is different for the mergers of type 2b (close pairs resolved into two galaxies by COMBO-17) because the individual galaxies making up the pair are complete only for $M _{*} \\geq10^{9}$ $M_{\\sun}$. As a result, we can only completely trace 1:3 major mergers of type 2b with total mass $\\ge 4.0 \\times 10^{9} M_{\\sun}$. If we impose this value to the afore-defined criterion of ($4 \\times M_{\\rm cut}/3$)}, it implies that a mass cut of $ M_{\\rm cut} \\ge 3.0 \\times 10^{9} M_{\\sun}$ is needed to ensure completeness for major mergers of type 2b. Similarly, we can only completely trace 1:9 minor mergers of type 2b with total mass $ \\ge 9.0 \\times 10^{9} M_{\\sun}$, which implies that a mass cut $M_{\\rm cut} \\ge 9.0 \\times 10^{9} M_{\\sun}$ is needed to completely trace all minor mergers of type 2b. Using a mass cut $M_{\\rm cut} \\ge 9.0 \\times 10^{9} M_{\\sun}$ for computing the merger fraction $f_{\\rm merge}$ would allow us to be complete for major and minor mergers of types 1, 2a, and 2b. However, it leads to very small number statistics and is not viable. We therefore explore the value of $f_{\\rm merge}$ for the less severe cut of $M_{\\rm cut} \\ge 3.0 \\times 10^{9} M_{\\sun}$, as well as for $M_{\\rm cut} \\ge 10^{9} M_{\\sun}$. The cut of $M_{\\rm cut} \\ge 3.0 \\times 10^{9} M_{\\sun}$ ensures that we are complete for major and minor mergers of type 1, 2a, but leaves us incomplete for mergers of type 2b. It is encouraging that both cuts yield similar values for the merger fraction $f_{\\rm merge}$, as illustrated in Table~\\ref{tcut}: the merger fraction is 0.023$\\pm$0.007 and 0.021$\\pm$0.007, respectively, for $M_{\\rm cut} \\ge 10^{9} M_{\\sun}$ and $M_{\\rm cut} \\ge 3.0 \\times 10^{9} M_{\\sun}$. Note that in computing the merger fraction, we only include the 20 distorted mergers listed in Table~\\ref{tmer}, and avoid the potential projection pairs of type 2a and 2b without signs of morphological distortions \\footnote{Even if we included all pairs of type 2a and 2b, the merger fraction would still have similar values (0.041$\\pm$0.01 and 0.033$\\pm$0.01) for both mass cuts (Table~\\ref{tcut}).}. While mergers may have played an important role in the evolution of cluster galaxies at earlier times (e.g., $z>2$), hierarchical models (e.g., Gottloeber et al 2001; Khochfar \\& Burkert 2001) predict that the merger fraction in dense clusters falls more steeply at $z<1$ than the field merger fraction. As a result, at $z<0.3$, the merger fraction for intermediate mass cluster galaxies is predicted to be quite low (typically below 5\\%). The low merger fraction among intermediate mass ($M = 10^{9}$ to a few $\\times 10^{10} M_{\\sun}$) galaxies in the A901/902 clusters is consistent with the latter prediction. How are the mergers distributed among major mergers, minor mergers, and ambiguous cases that could be either major or minor mergers? The results are shown in Columns 8--10 of Table ~\\ref{tcut}, based on the classification listed in Column 8 of Table ~\\ref{tmer}. The estimated fractions of likely major mergers, likely minor mergers, and ambiguous cases are 0.01$\\pm$0.004\\% (9/886), 0.006$\\pm$0.003\\% (5/886), and 0.007$\\pm$0.003\\% (6/886), respectively for $M_{\\rm cut} \\ge 10^{9} M_{\\sun}$. For $M_{\\rm cut} \\ge 3.0 \\times 10^{9} M_{\\sun}$, the corresponding fractions are 0.013$\\pm$ 0.005\\% (8/609), 0.008$\\pm$0.004\\% (5/609), and 0.0\\% (0/609), respectively. In the rest of this paper, we continue to work with a mass cut of $M_{\\rm cut} \\ge 10^{9} M_{\\sun}$, and the 20 distorted mergers (Table~\\ref{tmer}) applicable for this mass cut. However, where relevant, we cite many of our results for both of the mass cuts ($M_{\\rm cut} \\ge 10^{9} M_{\\sun}$ and $M_{\\rm cut} \\ge 3.0 \\times 10^{9} M_{\\sun}$) so that we can gauge the potential effect of incompleteness in tracing major and minor mergers. \\subsection{Frequency of mergers in A901/902 from CAS}\\label{scas1} The results of running CAS on the final classified sample of 886 intermediate-mass ($M_{*} \\geq 10^{9}$ $M_{\\sun}$) systems are shown in Table~\\ref{tcas} and Figures~\\ref{fcas1} and~\\ref{fcas2}. In Figure~\\ref{fcas1}, the 20 mergers of type 1, 2a, and 2b are plotted in different symbols. For Type 2b merger pairs, we plot the highest value of CAS $A$ found between the two galaxies in each system. Using the CAS merger criterion ($A>0.35$ and $A>S$) to identify mergers yields a merger fraction ($f_{\\rm CAS}$) in this sample of 18/886 or $0.02\\pm 0.006$. For the more conservative sample with a mass cut $M_{\\rm cut} \\ge 3 \\times 10^{9} M_{\\sun}$, $f_{\\rm CAS}$ is 7/609 or $0.011\\pm0.005$. When citing the error on $f_{\\rm CAS}$, we take the largest of either the Poisson error or the systematic error ($\\sigma_{\\rm CAS}$) in $f_{\\rm CAS}$ due to the systematic errors in CAS $A$ and $S$. The systematic error, $\\sigma_{\\rm CAS}$, is calculated by taking the upper and lower bounds of $f_{\\rm CAS}$ based on ($ A\\pm$ error in $A$) and ($S \\pm$ error in $S$). Specifically, these limits on $f_{\\rm CAS}$ are found by using the criteria of (($A \\pm$ error in $A$) $<$ ($S \\pm$ error in $S$)) and (($A \\pm$ error in $A$) $>$ 0.35). At first sight, the CAS-based merger fractions $f_{\\rm CAS}$ for the two mass cuts ($0.020\\pm0.006$ and $0.011\\pm0.005$ for $M_{\\rm cut} \\ge 10^{9} M_{\\sun}$ and $M_{\\rm cut} \\ge 3\\times 10^{9} M_{\\sun}$, respectively) are not widely different from the merger fraction $f_{\\rm merge}$ based on visual classification (0.023$\\pm$0.007 and 0.021$\\pm$0.007 for $M_{\\rm cut} \\ge 10^{9} M_{\\sun}$ and $M_{\\rm cut} \\ge 3\\times 10^{9} M_{\\sun}$, respectively; $\\S$Section~\\ref{sinte1}). However, the comparison of $f_{\\rm CAS}$ and $f_{\\rm merge}$ does not tell the whole story because the nature of the systems picked by the two methods can be quite different. The visual classes of the 18 systems, which satisfy the CAS criterion and are considered as mergers in the CAS system, are shown in Table~\\ref{tcas}. It turns out 7/18 ($39\\pm14\\%$) of these ``CAS mergers'' are visually classified as non-interacting systems. The results are illustrated in Table~\\ref{tcas} and Figure~\\ref{fcas1}. Figure~\\ref{fcas2} shows examples of these ``contaminants'': they tend to be dusty highly inclined systems and systems with low level asymmetries that seem to be caused by SF. It is also useful to ask what fraction of the 20 systems visually-identified as mergers satisfy the CAS criterion. We find that for $M_{\\rm cut} \\ge 10^{9} M_{\\sun}$, the CAS criterion only captures 11 of the 20 ($55\\pm16\\%$) of the visually-classified mergers. These results are illustrated in Figure~\\ref{fcas1}. Figure~\\ref{fcas2} shows examples of merging galaxies missed by CAS. The missed mergers include galaxies with fainter outer tidal features, double nuclei where CAS puts the center between the nuclei, and pairs of very close connected galaxies. It is also interesting to ask what percentage of the different Types of mergers (Types 1, 2a, and 2b) does the CAS merger criterion recover. Figure~\\ref{fcas1} also shows the mergers divided up by merger types 1, 2a and 2b. Of the 13 mergers of type 1, CAS recovers 9/13 or 69\\%$\\pm$18\\%. Of the 3 mergers of type 2a, CAS recovers 1/3 or 33\\%$\\pm$28\\%. Of the 4 mergers of type 2b, CAS recovers 1/4 or 25\\%$\\pm$22\\%. Since the CAS criterion is widely used to pick major mergers, it is also interesting to explore how well this criterion picks up the systems that we visually classified as major mergers (Table~\\ref{tmer}). We find that the CAS merger criterion picks up 6/9 (67\\%$\\pm20\\%$) of the systems classified as major mergers. It is also interesting to note that the CAS criterion recovers 2/5 (40\\%$\\pm23\\%$) and 4/6 (67\\%$\\pm23\\%$) of the systems classified respectively as minor mergers and ambiguous mergers. \\subsection{Distribution of mergers}\\label{sinte2} In order to define different regions of the A901a, A901b, and A902 clusters, we computed the projected number density $n$ (Figure~\\ref{fnumd}) for intermediate-mass ($M_{*} \\geq 10^{9} M_{\\sun}$) galaxies as a function of clustocentric radius $R$ by assuming a spherical distribution. Note here that each galaxy is assigned to the cluster closest to it, and $R$ is measured from the center of that cluster. We consider the cluster core to be at $R\\le$~0.25 Mpc, as this is the region where the projected number density $n$ rises very steeply (Figure~\\ref{fnumd}). The cluster virial radii are taken to be $\\sim$1.2 Mpc, based on estimates from the DM maps derived from gravitational lensing by Heymans \\etal (2008). Throughout this paper, we refer to the region at 0.25 Mpc~$$~1/10, as a function of environment, as characterized by the local overdensity ($\\delta^{\\rm G}$). The latter is calculated by smoothing the density of DM halos with a Gaussian of width 0.4 Mpc to take out the effect of individual galaxies. Typical values of $\\delta^{\\rm G}$ are $\\sim$10-100 for group overdensities, $\\sim$200 at the cluster virial radius, and $\\gtrsim$~1000 in the core of rich clusters. Typically, in the simulations, as field and group galaxies fall into a cluster along filaments, the coherent bulk flow enhances the galaxy density and causes galaxies to have small relative velocities, thus leading to a high probability for mergers at typical group overdensities. Closer to the cluster core, galaxies show large random motions rather than bulk flow motion, leading to a large galaxy velocity dispersion, and a sharp drop in the probability of mergers. In order to compare the data to simulation results, we use the number ($N_{\\rm merge}$) and fraction ($f_{\\rm merge}$) of mergers of mass ratio $\\ge 1/10$, which we computed in $\\S$Section~\\ref{sinte1}. The solid curve in Figure~\\ref{felco1} shows the merger fraction and number density predicted by the simulations as a function of overdensity. The three dashed lines in Figure~\\ref{felco1} show the estimated range in the observational number density ($n_{\\rm merge}$) and fraction ($f_{\\rm merge}$) of mergers in the three different regions (the core, the outer region, and the outskirt) of the A901/902 clusters, as defined in $\\S$Section~\\ref{sinte2}. The points at which the dashed lines cross or approach the solid curve tell us the typical overdensities at which we expect to find such merger fraction or merger number densities in the simulations. It can be seen that the low merger density seen in the core region of the cluster correspond to those expected at typical cluster core overdensities. On the other hand, the larger merger fraction we observe between the cluster core and outer region (0.25 Mpc~$< R \\le$~2 Mpc) is close to those seen in typical {\\it group overdensities}. Our results are in agreement with the above scenario, whereby the accretion of group and/or field galaxies along filaments, where they have low relative velocities and enhanced overdensities, leads to enhanced merger rates. We note that similar conclusions are reached if we perform the comparisons between data and simulations using the more conservative mass cut of $M_{*} \\geq 3 \\times 10^{9} M_{\\sun}$, discussed in $\\S$Section~\\ref{sinte1}. However, there are a couple of caveats when comparing our data with simulations. One caveat is that we directly compare the projected values of the number density or fraction of mergers from two dimensional observations, to the predicted number density of mergers from three dimensional simulations. Projection effects can introduce uncertainties in our observational estimate. The magnitude of the uncertainties due to projection effects depends on the detailed environment and will be investigated in the full STAGES supercluster simulation dataset (van Kampen \\etal 2009, in preparation). A further indirect evidence for group accretion stems from comparing semi-analytic galaxy catalogs to STAGES observations. Rhodes \\etal (in preparation) finds an overabundance in galaxies in A902 compared to its mass. This could be explained by two or more galaxy groups in projection, consistent with the idea of group accretion. \\subsection{Comparison with groups and clusters at different epochs}\\label{scomp1} We first recapitulate our results on the visually-based merger fraction ($f_{\\rm merge}$) in the A901/902 clusters ($\\S$Section~\\ref{sinte1}). Among intermediate-mass ($M_{*} \\geq10^{9} M_{\\sun}$) systems, we find that the fraction $f_{\\rm merge}$ of systems which show evidence of a recent or ongoing merger of mass ratio $ >$~1/10 is 0.023$\\pm$0.007 (Table~\\ref{tcut}). Most of these mergers have $M_{\\rm V} \\sim -19$ to $-22$ and $L \\le L^{*}$ (see Figure~\\ref{flumf}). We also estimated that the fraction of likely major mergers, likely minor mergers, and ambiguous cases to be 0.01$\\pm$0.004\\% (9/886), 0.006$\\pm$0.003\\% (5/886), and 0.007$\\pm$0.003\\% (6/886), respectively. Next, we compare our merger fraction in the A901/902 clusters with the published merger fraction in other clusters and group galaxies out to $z\\sim$~0.8, over the last 7 Gyr (Figure~\\ref{fcompa}). These comparisons are difficult to make as different studies apply different luminosity or mass cuts. Furthermore, some studies consider only major mergers, while others consider all interacting galaxies, which are likely candidates for both major and minor interactions. The variation in galaxy populations sampled at different epochs must also be kept in mind when comparing results at lower redshift with those out to $z\\sim$~0.8. Low redshift samples typically sample a small volume and therefore tend to host only a small number of the brightest and most massive systems. Conversely, higher redshift magnitude-limited samples will suffer from Malmquist bias and tend to under-represent the faint ($L < L^{*}$) galaxies. There have been several studies of galaxy mergers and interactions in intermediate redshift clusters (Lavery \\& Henry 1988; Lavery, Pierce, \\& Mclure 1992; Dressler \\etal 1994; Oemler, Dressler, \\& Butcher 1997; Couch \\etal 1998), but only few to date with high resolution $HST$ imaging. Couch \\etal (1998) used WFPC/WFPC2 observations and spectroscopy of two clusters at $z\\sim$0.3 and $\\sim0.4$, focusing on galaxies with $R\\sim$22.5 or $L < L^{*}$. In their study, they classified $\\sim$7/200 or $\\sim3.5\\%$ galaxies to be merging based on separations and visible distortions. The merger fractions of these two intermediate redshift clusters are plotted in Figure~\\ref{fcompa}. It is also interesting to note that the merging systems in these two intermediate redshift clusters tend to be blue and preferentially located in the outskirt of the clusters. In fact, in the Couch \\etal (1998) sample, $\\sim60\\%$ of the merging galaxies among $L 2.5 \\times 10^{13}$ $M_{\\sun}$) are major mergers at $z$ of 0.01--0.12. Most of the mergers involve two red sequence galaxies and they are located between 0.2 and 0.5 Mpc from the group center. A study by McGee \\etal (2008) also finds an enhancement in asymmetric bulge-dominated galaxies in groups, consistent with a higher probability for merging in the group environment. Additional evidence for mergers in groups is shown in a study of a supergroup at $z\\sim0.37$ by Tran \\etal (2008), who report dry dissipationless mergers and signatures thereof in three of four brightest group galaxies. A merger fraction of 6\\% was found by Zepf (1993) in Hickson compact groups at $z< 0.05$, among systems with luminosities $L \\leq L^{*}$. The merger fraction is significantly higher than that in SDSS groups. The increased merger fraction is expected since Hickson compact groups are different from loose groups: they have high number densities comparable to those in cluster cores, but low galaxy velocity dispersions. These two conditions provide an environment most favorable to strong tidal interactions and mergers, as argued in $\\S$Section~\\ref{sinte3}. It is not straightforward to draw conclusions about the evolution of the merger fraction in cluster galaxies over $z\\sim$~0.05--1.0 from the above studies because they sample different types of systems and different luminosity ranges. If we conservatively restrict ourselves to only studies of $L \\leq L^{*}$ cluster galaxies, then we have 4 data points over $z\\sim$~0.17--0.4, shown as solid filled circles in Figure~\\ref{fcompa}. The mean value of the data points allows for evolution by a factor of $\\sim$3.2, in the merger fraction of $L \\leq L^{*}$ cluster galaxies over $z\\sim$~0.17--0.4. However, if one uses the full range of merger fractions allowed by the error bars on the data points, then we can admit a wider spectrum of scenarios, ranging from no evolution to evolution by a factor of $\\sim$5 over $z\\sim$~0.17--0.4. Having additional deep, large-volume, high resolution studies, which are based on larger samples with smaller error bars would help to separate between these scenarios, and thereby test hierarchical models of galaxy evolution. As mentioned in $\\S$Section~\\ref{sinte1}, hierarchical models (e.g., Gottloeber et al 2001; Khochfar \\& Burkert 2001) predict that the merger fraction in dense clusters falls more steeply at $z<1$ than the field merger fraction, such that at $z<0.3$, the predicted merger fraction is quite low (typically below 5\\%) among intermediate mass cluster galaxies. The low merger fractions among intermediate mass ($10^{9}$ to a few $\\times 10^{10} M_{\\sun}$) or intermediate luminosities ($L < L^{*}$) galaxies in the A901/902 clusters and other low redshift clusters are consistent with these predictions, but we cannot yet test the predicted rate of evolution of the merger fraction in clusters with redshfit. \\subsection{Galaxies on the blue cloud and red sample}\\label{scolor1} We explore the properties of galaxies on the blue cloud and in the red sample among the final sample of 886 systems (20 mergers and 866 non-interacting galaxies; $\\S$Section~\\ref{smvc}). The red sample was defined in Wolf \\etal (2009) and contains both passively evolving spheroidal galaxies on the red sequence, as well as dusty red galaxies that lie between the red sequence and the blue cloud. Figure~\\ref{fcmd1} shows the rest-frame $U\\!-\\!V$ color plotted against stellar masses for galaxies of different visual classes: Mergers, Non-interacting Irr-1, and Non-interacting Symmetric ($\\S$Section~\\ref{smvc}). The visual classes of galaxies on the blue cloud and in the red sample are shown in Table~\\ref{tvccol}. As described in $\\S$Section~\\ref{sinte1}, we only consider the 20 distorted mergers listed in Table~\\ref{tmer}, and avoid the potential projection pairs without signs of morphological distortions. The 20 mergers are divided into 13 mergers of type 1, 3 mergers of type 2a, and 4 mergers of type 2b. For mergers of type 2b, which are resolved into two galaxies with separate COMBO-17 colors, we plot the average $U-V$ color of the galaxies in the pair. Out of the 886 visually classified systems in our sample, we find that 310/886 or $35\\%\\pm7\\%$ lie on the blue cloud. Out of the 20 visually classified mergers with distortions, 16/20 or $80\\%\\pm 18\\%$ lie on the blue cloud (Table~\\ref{tvccol}). Conversely, out of 866 non-interacting galaxies, 294/866 or $34\\%\\pm 7\\%$ are on the blue cloud. Thus, the fraction of mergers, which lie on the blue cloud ($f_{\\rm merg-BC}$) is over two times larger than the fraction of non-interacting galaxies ($f_{\\rm non-int-BC}$), which lie on the blue cloud. This implies that mergers and interacting galaxies are preferentially blue, compared to non-interacting galaxies . The observed higher fraction of blue galaxies among mergers compared to non-interacting galaxies may be caused by several factors. It may be due to enhanced levels of unobscured SF in mergers (see $\\S$Section~\\ref{ssfr1}), translating to bluer colors on average. It may also in part be the result of the mergers being part of accreted field and group galaxies ($\\S$Section~\\ref{sinte3}), which are bluer than the average cluster galaxy. It is also relevant to ask whether we are overestimating the fraction of blue galaxies among interacting systems due to the visibility timescale $t_{\\rm vis}$ of morphological distortions induced by interactions being longer in bluer galaxies than redder ones. While this is possible, it is non-trivial to correct for this effect because no direct unique relation exists between $t_{\\rm vis}$ and color. As discussed in $\\S~\\ref{sinte1}$, $t_{\\rm vis}$ depends on the mass ratio of an interaction as well as the gas content. A higher gas content may result in a longer $t_{\\rm vis}$ for certain gas distributions, but it can lead to either redder colors (e.g., enhanced level of dusty SF) or bluer colors (enhanced level of unobscured SF), compared to interacting systems. \\subsection{SF properties of interacting galaxies}\\label{ssfr1} This work uses SFRs based on UV data from COMBO-17 (Wolf \\etal 2004) and $Spitzer$ 24$\\micron$ imaging (Bell \\etal 2007). The unobscured SFR$_{\\rm UV}$ is derived using the 2800\\ \\AA \\ rest frame luminosity ($L_{\\rm UV}$ = 1.5$\\nu l_{\\nu,2800}$) as described in Bell \\etal (2005, 2007). The UV spectrum is dominated by continuum emission from massive stars and provides a good estimate of the integrated SFR from the younger stellar population in the wavelength range of 1216-3000\\ \\AA. \\ The SFR$_{\\rm IR}$ is derived using the 24$\\micron$ flux to construct the integrated IR luminosity ($L_{\\rm IR}$) over 8-1000$\\micron$ following the methods of Papovich \\& Bell (2002). The total SFR is derived using identical assumptions of Kennicutt (1998) from PEGASE assuming a 100 Myr old stellar population with constant SFR and a Chabrier (2003) IMF: \\begin{equation}\\label{sfrtot} {\\rm SFR}_{\\rm UV + IR} = 9.8 \\times 10^{-11}(L_{\\rm IR} + 2.2L_{\\rm UV}). \\end{equation} The factor of 2.2 on the UV luminosity accounts for light being emitted longward of 3000\\ \\AA \\ and shortward of 1216\\ \\AA \\ by young stars. The total SFR accounts for both the dust-reprocessed (IR) and unobscured (UV) SF. We work with our final sample of 886 classifiable bright massive ($M_{*} \\geq 10^{9} M_{\\sun}$) systems, and only consider here the 20 distorted mergers in Table~\\ref{tmer}. All of these systems have UV-based SFR from COMBO-17 observations. Of this sample, $\\sim$11\\% (94/886) were not observed with $Spitzer$, $\\sim$23\\% (206/886) were observed and detected at 24$\\micron$, while the rest had no detection at the $\\sim$4$\\sigma$ depth of 83 $\\mu$Jy. The UV-based SFR (SFR$_{\\rm UV}$) versus stellar mass is plotted in Figure~\\ref{fsfruv} for all 886 systems. The SFR$_{\\rm UV}$ ranges from $\\sim$0.01 to 14 $M_{\\sun}$ yr$^{-1}$. The UV-based SFR represents a lower limit to the total SFR for galaxies on the blue cloud and most star-forming galaxies on the red sample. However, for some old red galaxies, the SFR$_{\\rm UV}$ may overestimate the true SFR as the UV light from such systems may not trace massive OB stars, but rather low to intermediate mass stars. Figure~\\ref{fsfrtot} shows the UV+IR-based SFR (SFR$_{\\rm UV + IR}$), which ranges from $\\sim$0.2 to 9 $M_{\\sun}$ yr$^{-1}$. For the 206 galaxies that were observed and detected at 24$\\micron$, the implied UV+IR-based SFR is plotted as stars in the lower panel of Figure~\\ref{fsfrtot}. For the 586 galaxies that are observed but undetected with $Spitzer$, we use the 24$\\micron$ detection limit as an upper limit on the 24$\\micron$ flux. The corresponding upper limit on the UV+IR-based SFR is plotted as inverted triangles in the lower panel of Figure~\\ref{fsfrtot}, and included in the calculation of the average UV+IR-based SFR, plotted in the middle panel. In a cluster environment, the competition between processes that enhance the SFR and those that depress the SFR, ultimately determine the average SFR of cluster galaxies. The first class of processes are strong close interactions: tidal and mergers (e.g., Toomre \\& Toomre 1972), and harassment (e.g., Moore \\etal 1996), which refers to the cumulative effect of weak interactions. The second class of processes include ram pressure stripping of cold gas out of the galaxy by the hot ICM (e.g., Gunn \\& Gott 1972), and strangulation (e.g., Larson \\etal 1980; Balogh, Navarro \\& Morris 2000). For the few mergers (orange line) present, the average SFR$_{\\rm UV}$ is typically enhanced by an average modest factor of $\\sim$2 compared to the both Non-interacting Symmetric and Irr-1 galaxies (purple, green and black lines). Similarly, the UV+IR-based SFR (SFR$_{\\rm UV + IR}$; and Figure~\\ref{fsfrtot}) of merging galaxies is typically enhanced by only an average factor of $\\sim$1.5 compared to the Non-interacting Symmetric galaxies (purple line) and to all Non-interacting galaxies (i.e Symmetric + Irr-1; black line). We note that a similar modest enhancement in the average SFR, by a factor of 1.5--2 is also found in mergers in the field over $z\\sim$~0.24--0.80 by Jogee \\etal (2008, 2009). This modest enhancement is consistent with the theoretical predictions of di Matteo \\etal (2007; see their Figure 10), based on a recent statistical study of several hundred simulated galaxy collisions. Modest SFR enhancements are also seen in galaxy pair studies in the field (Barton \\etal 2000,2003; Lin \\etal 2004; Ellison \\etal 2008) and in mixed environments (Robaina \\etal 2009; Alonso \\etal 2004). While mergers in the A901/902 clusters enhance the SFR of individual galaxies, it is clear that they do not contribute much to the total SFR of the cluster. We compute the SFR density of mergers to non-interacting galaxies in the same volume of A901/902 by taking the ratio of the total SFR in each class. If we include only the 20 distorted mergers in Table~\\ref{tmer}, we find that the contribution of mergers to the SFR density of the clusters to be 10\\%. Alternatively, if we include all of the 36 visually classified mergers in Table~\\ref{tmer}, before accounting for false projection pairs, we find the contribution of mergers to the SFR density to be 15\\%. Thus, we find that mergers contribute only a small fraction (between 10\\% and 15\\%) of the total SFR density of the A901/902 clusters compared to non-interacting galaxies. The small contribution of mergers to the total cluster SFR density is likely due to the low number of mergers in A901/902, and the fact that these mergers only cause a modest SFR enhancement." }, "0910/0910.0883_arXiv.txt": { "abstract": "We describe two-dimensional gasdynamical computations of the X-ray emitting gas in the rotating elliptical galaxy NGC 4649 that indicate an inflow of $\\sim1$ $M_{\\odot}$ yr$^{-1}$ at every radius. Such a large instantaneous inflow cannot have persisted over a Hubble time. The central constant-entropy temperature peak recently observed in the innermost 150 parsecs is explained by compressive heating as gas flows toward the central massive black hole. Since the cooling time of this gas is only a few million years, NGC 4649 provides the most acutely concentrated known example of the cooling flow problem in which the time-integrated apparent mass that has flowed into the galactic core exceeds the total mass observed there. This paradox can be resolved by intermittent outflows of energy or mass driven by accretion energy released near the black hole. Inflowing gas is also required at intermediate kpc radii to explain the ellipticity of X-ray isophotes due to spin-up by mass ejected by stars that rotate with the galaxy and to explain local density and temperature profiles. We provide evidence that many luminous elliptical galaxies undergo similar inflow spin-up. A small turbulent viscosity is required in NGC 4649 to avoid forming large X-ray luminous disks that are not observed, but the turbulent pressure is small and does not interfere with mass determinations that assume hydrostatic equilibrium. ", "introduction": "X-ray isophotes of massive, rotating elliptical galaxies convey valuable information about the rotation of the hot, virialized interstellar gas, the interaction of this gas with gas recently ejected from evolving stars, and the sense of radial motion -- in or out -- of the hot gas. These issues are relevant to understanding the physical nature of the hot gas in galaxy groups and clusters in which gas loses energy by radiating X-rays but does not appear to cool to low temperatures as expected. In traditional cooling flows the radiating gas approximately maintains the local virial temperature as it flows slowly toward the center of the confining gravitational potential, then cools catastrophically near the center. The so-called cooling flow problem arises because there is insufficient spectroscopic evidence for cooling gas, previously cooled gas or recently formed stars. This paradox is apparent in massive galaxy clusters, galaxy groups and individual elliptical galaxies, i.e. the cooling flow problem is independent of scale. For this reason observations of X-ray luminous, relatively nearby elliptical galaxies can provide important clues to unravel the mysteries of cooling flows that do not seem to cool. The recent deep 81ks Chandra observation of the nearby X-ray luminous elliptical galaxy NGC 4649 by Humphrey et al. (2008) led to the detection of a massive central black hole ($M_{bh} = 3 \\times 10^{9}$ $M_{\\odot}$). This supermassive black hole was found for the first time directly from X-ray data by assuming that the hot gas is very nearly in hydrostatic equilibrium. The increased quality of the X-ray observations of NGC 4649 has motivated us to develop a dynamical model for the hot gas flow in this rotating galaxy. At any time the radial flow of hot gas must either be inward, outward, a combination of the two, or stationary. Many people think that cooling inflows cannot occur in elliptical galaxies or elsewhere because of the absence of X-ray spectral emission at low and intermediate temperatures (e.g. Xu et al. 2002). We showed for galactic-scale cooling flows that dust expelled from stars can cool the gas so fast that very little low-temperature thermal X-ray emission occurs (Mathews \\& Brighenti 2003a). Nevertheless, cooling inflows cannot dominate over time since the central mass accumulated would greatly exceed limits set by the velocity dispersion of central stars. Detailed gasdynamical models with {\\it ad hoc} heating sources have shown that the transition from cooling inflows to strong low-density wind outflows is very abrupt (Mathews \\& Brighenti 2003b). In the absence of extreme fine tuning, most of the X-ray luminous hot gas bound to elliptical galaxies cannot be flowing globally outward, although inhomogeneous outflowing regions of lower gas density (and X-ray emissivity) in jets or buoyant regions are not ruled out. If the gas is not flowing at all, continued enrichment by Type Ia supernovae for a few Gyrs will raise the iron abundance to several times that of the underlying stars, which is not generally observed (e.g. Humphrey \\& Buote 2006). One way to avoid this enrichment catastrophe is to recognize that some or most of the SNIa iron may cool radiatively in a way that cannot be observed, as discussed by Brighenti \\& Mathews (2005). However, the modest SNIa enrichment observed in the hot gas can be understood if much of the hot galactic gas flows radially inward, and the abundance can be even further reduced if more extended circumgalactic group-scale gas with relatively low metallicity flows inward past the sources of stellar enrichment. Circumgalactic gas is likely since Humphrey et al. (2006) derive a group-level virial mass for NGC 4649, $3.5 \\times 10^{13}$ $M_{\\odot}$. Several additional observations of NGC 4649 can be understood most easily in terms of gaseous inflow. First, the central gas temperature peak discovered by Humphrey et al. (2008) within a few 100 parsecs was predicted by cooling {\\it inflow} models as gas approaches a massive central black hole (Brighenti \\& Mathews 1999). While only a central gas pressure peak is sufficient to determine the black hole mass, a peak in the gas temperature is a signature of compressional heating as hot gas flows inward toward a strongly concentrated mass or small rotating disk. This compressive heating can occur even when the gas is losing energy by standard radiative losses. Additional support for subsonic compressional heating during inflow follows from the observation that the entropy of hot gas closest to the black hole is lower than in any other region in NGC 4649 so it has not been recently heated by the central AGN. Another important signature of subsonically inflowing gas seen in NGC 4649 and other similar elliptical galaxies is a central flattening of X-ray isophotes on kpc scales. Elliptical X-ray isophotes on galactic scales have different implications than on larger scales which have been used to infer flattening of the dark halo potential (Buote et al. 2002). Isophotal flattening in excess of that required by the underlying stellar potential can be interpreted as a spinning up of hot inflowing gas expected when angular momentum is exchanged with gas recently ejected from old, mass-losing red giant stars that participate in the galactic rotation. We review below the recent compilation of Chandra isophotes of elliptical galaxies by Diehl \\& Statler (2007) and show that they generally resemble NGC 4649 in central flattening and implied inflow spin-up. NGC 4649 is an ideal candidate galaxy to study the influence of rotation on the X-ray isophotes because of its remarkably high rotation rate for a massive, cored E galaxy, its proximity, and our good fortune to have received at least one moderately deep Chandra observation. In the following discussion we review the implications of recent Chandra images of NGC 4649 and present gasdynamical models of increasing complexity. Unlike our previous studies of rotating cooling flows (Brighenti \\& Mathews 1996, 2000), we include the important contribution of inflowing and non-rotating circumgalactic gas from a group-scale halo. In addition, since the central X-ray isophotal flattening observed in NGC 4649 and other massive ellipticals is less than that expected from the conservation of angular momentum, we invoke an {\\it ad hoc} hot gas turbulence that can transport angular momentum away from the galactic spin axis. By combining turbulence and circumgalactic inflow it is easy to explain the isophotal flattening observed in NGC 4649. Moreover, the turbulent viscosity that is consistent with the X-ray isophotes of NGC 4649 generates a turbulent pressure that is much less than the gas pressure. Therefore, the subsonic turbulent activity required to reduce the angular momentum does not significantly degrade the total mass determinations of Humphrey et al. (2008) based on assuming hydrostatic equilibrium. We assume that the mild turbulence we require is produced by energy released near the central black hole or active galactic nucleus (AGN). Our treatment here is done under the assumption that most of the {\\it observed} X-ray emission in NGC 4649 can be explained with inflowing gas. We do not propose that this is the complete solution to the gas flow problem in this or other similar galaxies, but only demonstrate that the X-ray observations are consistent with an {\\it apparent} global inflow at the present time. To avoid unobserved catastrophic radiative cooling and a huge mass concentration in the galactic core, we must also assume that there is a (possibly buoyant) return mass outflow that is too intermittent, too hot, or too cunningly disguised to contribute to Chandra observations (Mathews \\& Brighenti 2004; Brighenti \\& Mathews 2006). ", "conclusions": "We have shown that the X-ray emitting gas associated with the bright elliptical galaxy NGC 4649 has density and temperature profiles that can be understood as a rotating, radiatively cooling inflow with a mild turbulent viscosity. Evidence of inflow is present at every radius in NGC 4649. On the largest scales, a radiatively cooling inflow of circumgalactic gas is required to dilute the hot gas iron abundance acquired from supernovae down to observed values, assuming that the iron ejected by Type Ia supernovae does not cool by radiative losses before it merges with the ambient hot gas. The relatively high gas density and temperature observed in NGC 4649 beyond $\\sim$10 kpc can also be explained with inflowing circumgalactic gas. On intermediate galactic scales, observations of the X-ray ellipticity in NGC 4649 and other bright elliptical galaxies show that the hot gas is spun up by mass ejected from evolving stars that rotate collectively with the galaxy. Inflow spin-up is seen in the increasing X-ray isophotal ellipticity with decreasing galactic radius until it exceeds the ellipticity of the local stars near the galactic center. However, for NGC 4649, which rotates unusually fast for a massive boxy elliptical, a small turbulent viscosity is required to avoid forming multi-kpc X-ray disks that are not observed. The shallow negative temperature gradient inside $\\sim1$ kpc in NGC 4649 is X-ray evidence of an inflow that compresses toward a sub-kpc disk. The observed X-ray isophotes can be matched with a turbulent viscosity for which the corresponding turbulent pressure is much less than the gas pressure, so the integrated mass profiles found by assuming hydrostatic equilibrium (Humphrey et al. 2008) are unaffected. This alleviates the concern expressed by Diehl \\& Statler (2007) that the hot gas in elliptical galaxies is very far from hydrostatic equilibrium. Galactic mass determinations based on X-ray thermal emission may fail for some elliptical galaxies where non-thermal pressure dominates the galactic core, such as NGC 4636 (Brighenti \\& Mathews 1997; Baldi et al. 2009), but this problem does not seem to occur in NGC 4649. On the smallest scales observed with Chandra, within about 100 parsecs, the central temperature peak in NGC 4649 discovered by Humphrey et al. (2008) is a natural consequence of subsonic cooling inflow. In spite of ongoing radiative losses, the central gas temperature increases due to (nearly adiabatic) compression as gas approaches the central black hole or the small disk around it. The kpc-scale negative temperature gradient in NGC 4649 formed as gas compresses toward a small disk is not caused by recent AGN heating since the gas entropy in this region increases with galactic radius. This type of central inflow is also consistent with nuclear disks of cooled dusty gas having radii of a few hundred parsecs commonly observed in the cores of luminous elliptical galaxies. NGC 4649 represents the most concentrated known example of the cooling flow problem. The observed density and temperature profiles can be explained at every radius with a radiatively cooling inflow. At the smallest observable radius in NGC 4649, about 150 pc, the times for cooling and flow to the core are only a few million years. For this reason it will be very difficult to devise heating scenarios that maintain the gas at rest, perfectly mimicking a cooling inflow. No ideally heated model in which the hot gas in NGC 4649 remains stationary can account for commonly observed X-ray ellipticity profiles $\\epsilon_X(r)$ that require inflow spin-up. The current mass inflow rate in our successful calculations is $\\sim1$ $M_{\\odot}$ yr$^{-1}$, but the stellar mass loss rate was certainly several times larger in the past. Consequently, if this flow continued over a Hubble time, the total mass of cooled gas accumulated in the galactic core of NGC 4649 would exceed the mass of the black hole and stars within $\\sim1$ kpc by factors of 3 - 10. Therefore, the mass inflow indicated by the central temperature peak must be removed from within $\\sim150$ pc at a (time-averaged) rate comparable to $\\sim1$ $M_{\\odot}$ yr$^{-1}$ and transported out to a large radius without interfering with the X-ray appearance of NGC 4649 which resembles a traditional cooling flow. This mass outflow may occur in the bipolar jets observed in NGC 4649 and many other bright ellipticals. Alternatively, nuclear hot gas may become buoyant by intermittent local heating or by cosmic rays and flow out subsonically upstream in the cooling flow (e.g. Mathews \\& Brighenti 2008). If the buoyant gas density is only slightly less than that of the ambient gas, it may not significantly disturb the radial gas density and temperature profiles set by the inflowing gas. Alternatively, if the density of buoyant (or jet) gas were considerably lower, its X-ray emission would be less easily observed against the brighter emission from denser inflowing gas. \\vskip.1in" }, "0910/0910.5142_arXiv.txt": { "abstract": "Cosmology is operating now on a well established and tightly constraining empirical basis. The relativistic $\\Lambda$CDM hot big bang theory is consistent with all the present tests; it has become the benchmark. But the many open issues in this subject make it reasonable to expect that a more accurate cosmology will have more interesting physics in the invisible sector of the universe, and maybe also in the visible part. ", "introduction": "\\begin{figure}[b] \\includegraphics[angle=90,height=.33 \\textheight]{cmb_spectrum.pdf} \\caption{Energy spectrum of the cosmic microwave background, from Alan Kogut.} \\end{figure} Lest the exciting debates over the great open issues of cosmology cause us to forget, I offer in Figure~1 and Table~1 reminders that we have a firm and substantial empirical basis for our subject. The figure shows the spectrum of the cosmic microwave radiation that nearly uniformly fills space: energy density as a function of frequency. The measurements are diverse and well cross-checked, from the ground, balloons, rocket and satellite missions, and indirectly from spin temperatures derived from interstellar molecular absorption lines. Within the measurement uncertainties, which over much of the range of frequencies in this figure are smaller than the symbols, this wealth of data fits the Planck thermal spectrum at temperature $T=2.725$~K plotted as the solid curve. We know the universe as it is now is close to transparent at the frequencies plotted here, because active galaxies at redshifts exceeding unity are detected at these frequencies. Relaxation to this thermal spectrum thus had to have happened earlier in the expansion of the universe, when the density of absorbing matter was great enough make the universe opaque across the Hubble length then. That is, the figure presents us with almost tangible evidence that the universe really did evolve from a very different state. We should take a moment to admire this wonderfully simple and deeply significant phenomenon. The figure also shows that expansion and cooling to the present transparent state of the universe had to have preserved the thermal radiation spectrum formed when the universe was opaque. Sufficient conditions include standard local physics and a metric description of spacetime. It does not require general relativity theory. It does require that, despite the distinctly inhomogeneous distribution of matter in galaxies and concentrations of galaxies, spacetime is close enough to homogeneous and isotropic not to have produced a sigificant spread of redshifts along different lines of sight (as could have happened in a homogeneous anisotropic universe, for example). Space has to be close to what Einstein envisioned in 1917. The $\\Lambda$CDM cosmology adopts Einstein's near homogeneous space, his general relativity theory, and his cosmological constant, and it adds the general expansion first discussed by Friedmann and Lema\\^\\i tre, Tolman's thermal radiation, Gamow's thermonuclear element formation, and the more recent notions of nonbaryonic dark matter and adiabatic near scale-invariant Gaussian initial departures from homogeneity. The application of general relativity is a long extrapolation from its precision tests, but it was natural to try this theory first. Other of these assumptions were not so obvious first guesses; they were adopted because they were seen to offer promising fits to the improving tests. We would be poor theorists indeed if $\\Lambda$CDM did not at least approximate what is now measured. But this cosmology has gone beyond that; it is predictive. \\begin{table}[t] \\caption{Cosmological Tests} \\centering \\begin{tabular}{lllll} \\hline \\multicolumn{2}{l}{1. fossil CMB radiation} & \\multicolumn{2}{l}{6. cosmic mass density} \\\\ & (a) energy spectrum && (a) galaxy peculiar velocities \\\\ & (b) temperature acoustic oscillation && (b) gravitational lensing \\\\ & (c) temperature-polarization cross spectrum & \\multicolumn{2}{l}{7. large-scale structure} \\\\ & (d) integral Sach-Wolfe effect && (a) baryon acoustic oscillation \\\\ \\multicolumn{2}{l} {2. light element abundances} && (b) galaxy N-point functions \\\\ & (a) deuterium in Ly$\\alpha$ absorbers at redshift $\\sim 3$ & \\multicolumn{2}{l}{8. clusters of galaxies}\\\\ & (b) helium in dwarf low-metallicity galaxies && (a) mass function as a function of mass \\& redshift \\\\ & (c) observed baryon budget && (b) baryon mass fraction \\\\ & (d) count of neutrino families & \\multicolumn{2}{l}{9. small-scale structure}\\\\ \\multicolumn{2}{l}{3. redshift-magnitude relation} && (a) galaxy structure and evolution \\\\ \\multicolumn{2}{l}{4. stellar evolution \\& radioactive decay ages} && (b) Lyman-$\\alpha$ forest\\\\ \\multicolumn{2}{l}{5. distance scale}\\\\ & (a) distance ladders\\\\ & (b) gravitational lensing\\\\ \\hline \\end{tabular} \\end{table} This situation is illustrated in Table~1 (adapted from the book, {\\it Finding the Big Bang} \\cite{FTBB}; details are there and in other reviews of the state of the cosmological tests). Each measurement named in the table is an independent (or close to it) probe of the universe. Each is capable of falsifying $\\Lambda$CDM. This cosmology passes the tests so far. The measurements are difficult, and issues of interpretation and systematic errors are well worth continued close examination. But the diversity of these probes, and the consistency of independent constraints on the relevant parameters in $\\Lambda$CDM, make a close to compelling case that we have not been seriously misled. That certainly does not mean the $\\Lambda$CDM cosmology is the final word on the structure and physics of the universe that will be needed to interpret coming generations of measurements. But we can be sure that if $\\Lambda$CDM is replaced by something better the new theory will predict a universe that looks much like $\\Lambda$CDM, because that is what is observed. \\subsection{Precision Cosmology and Accurate Cosmology} We should take another moment to admire the abundance of probes of the large-scale nature of the universe represented in Table~1, and to recall the importance of the diversity of these measurements. It has been said that we are moving toward precision cosmology, which is correct but misses the point. A digital scale may read out the weight of an object to many significant figures, in a precise measurement. But if the scale is not well calibrated the measurement may not be very accurate. That is, accuracy is what remains of precision after discounting systematic errors. And our goal is an accurate cosmology. In the pursuit of precision cosmology it is natural to concentrate on the relatively few observations that lend themselves to the most precise measurements. This is good science, provided one bears in mind that if the precision measurements are not more numerous than the parameters that may be adjusted to fit them we are getting a cosmology of dubious accuracy. I offer one example, the properties of galaxies. The mass distributions in the outer parts of galaxies were an apparent anomaly within the old cosmology, and a hint of something new: dark matter. Recent research proceeds in the opposite direction: take $\\Lambda$CDM, dark matter and all, as given, to be used as the basis for analyses of how galaxies acquired their observed properties. Since galaxies are complicated --- our understanding of star formation and its impact on the diffuse baryons is particularly schematic --- the program must add parameters that are physically well motivated but adjustable because we don't understand how the physics actually is expressed. The freedom of adjustment makes it difficult to judge the significance of the notable advances in this program. The pursuit of more accurate cosmology may be aided by a return to the earlier direction. There are apparent anomalies in the properties of galaxies in terms of what might have been expected in a straightforward reading of numerical simulations of structure formation in $\\Lambda$CDM. Instead of asking how the parameters needed to describe the histories of the baryons may be adjusted to remedy this situation consider instead whether apparent anomalies might be hints to better fundamental physics. It is not difficult to invent alternative physics that fits the tests on the scale of the Hubble length about as well as $\\Lambda$CDM, but on the scale of galaxies does something different, interesting, and just possibly better. An example of an alternative theory on the scale of galaxies is discussed next; others are in \\cite{Khoury} -- \\cite{Carroll et al.}. ", "conclusions": "Cosmology has changed from the mode of operation the older of us remember without nostalgia. We are now tightly constrained by a rich observational basis that makes it exceedingly delicate --- though certainly not impossible --- to find and demonstrate viability of alternatives to the $\\Lambda$CDM cosmology. We each have chosen a mode of research appropriate for this relatively new situation in cosmology. It is worth listing strategies, to remind ourselves and funding agencies that we do have choices. It is entirely reasonable to take the conservative position that $\\Lambda$CDM, maybe with slow evolution of the dark energy density and fine tuning of initial conditions, includes all the physics that is going to be relevant to analyses of phenomenological cosmology. Then the goal becomes to understand how this given physics is expressed in the processes that make the universe we observe, from the Hubble length down to galaxies and stars and planets, and back in time to light element production and maybe baryogenesis. This strategy has proven its worth by defining a host of productive research problems. If $\\Lambda$CDM does differ from reality enough to matter it means this conservative strategy will reveal it through identification of phenomena that cannot be reconciled with the cosmology. But beware of precision cosmology, because the observations that lend themselves precision measurements reduce the tests in Table~1 to a number perilously close to the parameters we are willing to adjust in $\\Lambda$CDM. (Currently discussed additions to the more traditional cosmological parameters include evolution and maybe spatial gradients of the dark energy density, a soup\\c{c}on of annihilating or collisional or warm dark matter, or maybe even a new long-range interaction in the dark sector. A competent theorist likely could offer quite a few other ways to save the phenomena within $\\Lambda$CDM.) I expect that the better strategy for finding a more accurate cosmology will be to examine all measurements, however inaccurate, that probe different aspects of the universe and may challenge standard ideas. A more proactive strategy for discovering a possibly more accurate cosmology is to search for apparent anomalies in $\\Lambda$CDM, situations that seem particularly challenging for this theory. My list commences with a pronounced difference between the physics of the visible and invisible universes. Physics in the visible sector is wonderfully simple, in the appropriate sense, but its expression is spectacularly complicated. Physics in the invisible sector is so exceedingly simple that its expression is simple: the dark matter just piles up in halos and the dark energy sits there or maybe slowly evolves. It is wishful thinking, but also sensible, to suspect that this apparently anomalous situation is telling us that we have settled for the simplest physics we can get away with at the present crude levels of probes of this sector. It presents no hints of phenomena to look for; I read it an invitation to look broadly. Other apparent anomalies may offer more direct hints of better physics. Examples include preliminary results from underground dark matter particle detectors and the search for signatures of dark matter annihilation. At the time of writing the experimental/observational situation is confused, maybe in part a result of systematic errors in difficult measurements, maybe in part because we are seeing hints to something new. Other examples are the more curious properties of galaxies. How did galaxies arrive at their scaling relations that show such little regard for environment? Why do the voids defined by normal galaxies look so very empty? Why do most of the nearest spiral galaxies edge up to the very empty Local Void? How did massive black holes in active galactic nuclei form so early in the expansion of the universe? Why is large-scale structure in the galaxy distribution so very large? If such curiosities are expressions of standard physics then understanding how they came about will teach us something of value. And maybe some are hints to better physics, which would be a good thing too. Research in the proactive direction may accept the standard physics of the visible universe and seek something more interesting in the poorly explored invisible part. As one sees in these Proceedings, there is no shortage of ideas here. The adoption of standard visible sector physics is conservative, but maybe overly so. For example, the precision tests of general relativity are on scales some fifteen orders of magnitude smaller than the Hubble length. It is sensible to continue to question this enormous extrapolation, to consider modifications of gravity physics. This strategy has led to viable alternatives that to date seem to me to be less logically compelling than general relativity with $\\Lambda$CDM. This is not a very convincing argument, but I expect it will lead most in the community to continue to work within the physics of $\\Lambda$CDM unless/until some new phenomenology or really attractive idea drives us from it. I refer to Wigner's \\cite{Wigner} essay, {\\it The Unreasonable Effectiveness of Mathematics in the Natural Sciences}, for the arguments that lead me to expect that the community will not arrive at the one true cosmology, if such a thing exists. We will instead end up with the best approximation allowed by the limited ability to observe. We cannot rule out the possibility that something will turn up that invalidates our present cosmology, but from all experience that is exceedingly unlikely. Much easier to imagine is fine tuning to fit new measurements will lead to a more accurate cosmology that contains elements of $\\Lambda$CDM along with new and interesting adjustments." }, "0910/0910.2554_arXiv.txt": { "abstract": "% The evolution of galaxy merger rates and its impact on galaxy properties have been studied intensively over the last decade. It becomes clear now that various types of mergers, i.e. gas-rich (wet), gas-poor (dry), or mixed mergers, affect the merger products in different ways. The epoch when each type of merger dominates also differs. In this talk, I review the recent progress on the measurements of galaxy merger rates out to $z \\sim 3$ and the level of interaction-triggered star formation using large samples from various redshift surveys. These results provide insights to the importance of mergers in the mass assembly history of galaxies and in the evolution of galaxy properties. I also present new results in characterizing the environment of galaxy mergers, and discuss their implications in the built up of red-sequence galaxies. ", "introduction": "% Galaxy mergers have long been suggested to be associated with a variety of observational phenomena, including Ultra-Luminous Infrared Galaxies (ULIRGs), Quasars, post-starburst galaxies, Active Galactic Nuclei (AGN), and Submillimeter Galaxies (SMGs), etc. In certain galaxy evolution models, these objects may actually correspond to different phases during galaxy interactions \\citep{hop06}. Galaxy mergers have significant impact on galaxy morphology, kinematics, stellar masses, and star formation histories of galaxies, as well as the mass of its central black holes. Recent studies on the galaxy luminosity function (LF) and stellar masses function (SMF) reveal that the number densities and stellar mass densities of galaxies in the red sequence have been increased by a factor of 2 to 3 since $z \\sim 1$ while that of blue forming galaxies remain similar over the same period \\citep{fab07,bel07}, suggesting a migration of galaxies from the blue cloud to the red sequence. One plausible mechanism responsible for this transformation is through merging of two blue galaxies, the so-called 'wet mergers' (gas-rich mergers) followed by subsequent mergers among re-sequence galaxies, the so-called 'dry mergers' (gas-poor mergers) \\citep{fab07}. Therefore in order to better understand how the present-day massive galaxies are assembled, and how the galaxies have been evolved, we need to have improved constraints on the abundance of mergers and where they occur. There are three aspects regarding galaxy interactions that I focus on in this talk: the absolute galaxy merger rates, the influence of mergers in triggering star formation, and the environment of merging galaxies. ", "conclusions": "I have argued that studies of merger rates, effectiveness of the tidally-triggered star formation, and the host environment of mergers are the three keys to pin down the role of mergers in the evolution history of galaxies. Despite that our understanding of the above issues have been dramatically improved over the last few years, there remains several pieces of details missing in the whole picture. For example, under what kind of conditions will wet mergers lead to the quench of star formation and hence become K+A galaxies? What is the outcome of mixed mergers? What is the merger rate beyond redshift one? How well can we separate the contributions of minor mergers from major mergers? And finally, what is the role of AGN during galaxy encounters? Both detailed studies of individual interacting systems and statistical results based on future larger and deeper surveys of merging galaxies, together with improved numerical modeling that incorporate more sophisticated physics will provide essential knowledge to resolve the aforementioned issues." }, "0910/0910.3691_arXiv.txt": { "abstract": "While feedback is important in theoretical models, we do not really know if it works in reality. Feedback from jets appears to be sufficient to keep the cooling flows in clusters from cooling too much and it may be sufficient to regulate black hole growth in dominant cluster galaxies. Only about 10\\% of all quasars, however, have powerful radio jets, so jet-related feedback cannot be generic. The outflows could potentially be a more common form of AGN feedback, but measuring mass and energy outflow rates is a challenging task, the main unknown being the location and geometry of the absorbing medium. Using a novel technique, we made first such measurement in NGC 4051 using XMM data and found the mass and energy outflow rates to be 4 to 5 orders of magnitude {\\it below} those required for efficient feedback. To test whether the outflow velocity in NGC 4051 is unusually low, we compared the ratio of outflow velocity to escape velocity in a sample of AGNs and found it to be generally less than one. It is thus possible that in most Seyferts the feedback is not sufficient and may not be necessary. ", "introduction": " ", "conclusions": "" }, "0910/0910.1007_arXiv.txt": { "abstract": "{} % {The co-evolution of host galaxies and the active black holes which reside in their centre is one of the most important topics in modern observational cosmology. Here we present a study of the properties of obscured Active Galactic Nuclei (AGN) detected in the CDFS 1Ms observation and their host galaxies.} {We limited the analysis to the MUSIC area, for which deep K-band observations obtained with ISAAC@VLT are available, ensuring accurate identifications of the counterparts of the X--ray sources as well as reliable determination of photometric redshifts and galaxy parameters, such as stellar masses and star formation rates. In particular, we: 1) refined the X-ray/infrared/optical association of 179 sources in the MUSIC area detected in the Chandra observation; 2) studied the host galaxies observed and rest frame colors and properties.} {We found that X--ray selected (L$_X\\gs10^{42}$ erg s$^{-1}$) AGN show Spitzer colors consistent with both AGN and starburst dominated infrared continuum; the latter would not have been selected as AGN from infrared diagnostics. The host galaxies of X--ray selected obscured AGN are all massive (M$_*>10^{10}$ M$\\odot$) and, in 50\\% of the cases, are also actively forming stars (1/SSFR$<$ t$_{Hubble}$) in dusty environments. The median L/LEdd value of the active nucleus is between 2\\% and 10\\% depending on the assumed M$_{BH}$/M$_{*}$ ratio. % Finally, we found that the X--ray selected AGN fraction increases with the stellar mass up to a value of $\\sim30$\\% at z$>1$ and M$_*>3\\times10^{11}$ M$\\odot$, a fraction significantly higher than in the local Universe for AGN of similar luminosities.} {} ", "introduction": "Galaxy interactions, and more in general the large scale structure (LSS) galaxy environment, are thought to play a major role in regulating both star-formation and accretion onto nuclear super-massive black holes (SMBHs). This implies that the full understanding of galaxy evolution requires a good knowledge of the SMBH census through cosmic time. In particular, feedback between the central SMBHs in their active phases and the interstellar medium is likely to affect strongly the evolution of their host galaxies. A short, powerful but highly obscured growth phase of both SMBHs and their host galaxies is predicted by many models for the co-evolution of galaxies and AGN (Silk \\& Rees 1998, Fabian 1999, Granato et al. 2004, Di Matteo et al. 2005, Menci et al. 2008). This phase ends when strong AGN winds and shocks heat the interstellar medium, blowing away the dust and gas and inhibiting further star-formation in the AGN host galaxies. According to this ``evolutionary sequence'' (e.g. Hopkins et al. 2008), highly obscured AGN should be associated to young galaxies in the process of assembling most of their stellar mass through significant episodes of star formation. On the contrary, unobscured AGN should be associated to galaxies with low or absent episodes of star-formation, given that most of the gas and dust responsible for the star formation has been blown away by the effect of AGN feedback. Many observational evidences (see Alexander et al. 2005, Page et al. 2004, Stevens et al. 2006) and theoretical arguments (Menci et al. 2008 and references therein) in favor of the evolutionary sequence do exist. These results challenge our 20-years old AGN view, in which the differences we see in different classes of sources - especially between \"obscured\" and \"unobscured\" ones - are simply due to orientations effects (Antonucci et al. 1985, Antonucci 2003, Urry \\& Padovani 1995). \\\\ A correct and complete identification of obscured and highly obscured AGN at high redshift is therefore crucial because they may represent the first, still little explored, phase of the common growth of both SMBHs and their host galaxies. Moderately and highly obscured AGN at z$\\sim2$, where most of the accretion and star-formation processes are on-going, offer an ideal tool for a direct test of feedback models. In these objects the nuclear light is completely blocked or strongly reduced and does not overshine the galaxy optical and near-infrared light. This gives us the possibility to study host galaxy morphologies, colors and spectral energy distributions without the difficulty of disentangling star-light from nuclear light. We can therefore study galaxy properties during active phases, e.g. {\\it when nuclear feedback should be in action}. Obscured AGN at cosmological distances are however usually faint in the observed optical bands, because the UV rest frame is strongly reduced by dust extinction. On the other hand, moderately obscured AGN (or Compton thin, $10^{22}<$N$_H<10^{24}$ cm$^{-2}$) are common in X-ray images (Bauer et al. 2004, Comastri \\& Brusa 2008 and reference therein), making up $\\sim$50\\% of the full X-ray population at fluxes $<10^{-14}$ \\cgs (Gilli, Comastri \\& Hasinger 2007) once the contribution from the normal starforming galaxy population is removed (Ranalli et al. 2005). Compton thick AGN (CT, N$_H\\gs10^{24}$ cm$^{-2}$) are faint also in the X-ray band, because photoelectric absorption and Compton scattering strongly reduce the X-ray flux up to $>10$ keV, and over the entire X--ray range if N$_H\\gs10^{25}$ cm$^{-2}$. Indeed, only a dozen CT AGN is present in the deepest images of the X-ray sky, the Chandra Deep Fields (Norman et al. 2002, Tozzi et al. 2006, Georgantopoulos et al. 2008). Identification of the correct counterparts of obscured AGN at cosmological redshift is not trivial. These objects are faint in the optical band, because 1) the intrinsic AGN emission is absorbed by the surrounding material, and 2) the host galaxy light is strongly reduced by cosmological dimming (L$^*$galaxies at z=1-2 have R$\\sim24-25$). The probability to find by chance a galaxy of these magnitudes in the Chandra error boxes is not negligible. The identification process is made easier by using deep near infrared images, because the surface density of these sources is smaller than that of optical ones, bringing the probability of finding a galaxy by chance in the Chandra error boxes to comfortably small values. Moreover, the K-band flux is more tightly correlated with the X-ray flux than the optical (obscured) one. For this reason, we have reanalyzed the identifications of the Chandra Deep Field South (CDFS) 1 Ms sources (Alexander et al. 2003) in the area covered by sensitive K band and IRAC 3.6$\\mu$m and 4.5$\\mu$m observations (the GOODS MUSIC area, Grazian et al. 2006). The excellent multiband photometry available in this area allows the determination of a reliable photometric redshift for the faint sources not reachable by optical spectroscopy. Once identified the counterparts of the X-ray sources and their redshift, it is possible to proceed to a detailed study of the properties of their host galaxies. We present here a study of the properties of the host galaxies of X-ray obscured AGN, such as their morphology and close environment, optical and infrared rest frame colors, mid infrared to optical spectral energy distributions. Galaxy masses and star-formation rates are derived from the fit of galaxy templates to the optical to near infrared spectral energy distributions. We finally compare the masses and star-formation rates of obscured AGN host galaxies to those of inactive galaxies in the field selected in both optical and infrared bands. The paper is organized as follows: Sect. 2 presents the CDFS datasets, the X-ray to optical/infrared association and the obscured AGN sample. Section 3 presents the observed frame colors of the obscured AGN sample. Section 4 presents the host galaxies properties (masses and star-formation rates, SFRs ) of the sample and our estimates of the fraction of AGN in mass selected samples. Section 5 presents a a discussion of the results and Section 6 outlines a summary of the most important points. A cosmology with $H_0=70$ km s$^{-1}$ Mpc$^{-1}$, $\\Omega_M$=0.3, $\\Omega_{\\Lambda}=0.7$ is adopted throughout (Spergel et al. 2003). ", "conclusions": "We presented a new catalog of the counterparts of the 179 extragalactic X--ray sources detected in the 1Ms Chandra observation of the MUSIC/CDFS/GOODS field and an extensive analysis of the host galaxies properties of obscured AGN. \\\\ We quantified the bias in the determination of the counterparts of X--ray selected sources when the match is limited to optical catalogs only, with respect to the combined use of optical and near infrared (deep K--band and IRAC) data. We estimate that the fraction of misidentified X--ray sources previously reported in the literature is of the order of $\\sim 6\\%$, and rises up to $\\sim14$\\% when optically faint ($z>24$) sources are considered (see Figure 1); the use of an optically-based catalog biases the identification against the most extreme, obscured sources therefore preventing the exact knowledge of the multiwavelength properties of the X-ray counterparts. \\\\ \\par\\noindent In order to study the host-galaxies of obscured AGN, we defined a sample of 116 ``bona fide'' obscured AGN, by selecting sources without broad lines in the optical spectra and with small optical nuclear emission with respect to the host galaxy optical emission (see Figure 3). From eyeball inspection of the host galaxy morphology, we found a variety of cases, and a disturbed morphology (due to activity/merging/star formation) in more than half of the sub-sample for which a morphological classification could be made (see Figure 4).\\\\ \\par\\noindent We investigated the optical to infrared colors of these obscured AGN. The most striking result is that half of the X--ray selected obscured AGN in the redshift range z=1.5-4.0 have a 8.0$\\mu$m to 4.5$\\mu$m flux ratio $<2$ and according to Pope et al. (2008) would have been classified as ``star-formation'' dominated objects (see Figure 6); 50\\% of them in addition have 24.0$\\mu$m to 8.0$\\mu$m flux ratio $<5$, where the original Pope et al. (2008) diagram is almost empty. Moreover, previous analysis based on Chandra stacking analysis of sources with high MIR/O (Daddi et al. 2007a, Fiore et al. 2008, 2009) claimed a large contribution ($\\geq80$\\%) of heavily obscured (Compton Thick) sources among the stacked population (but see also discussion in Donley et al. 2008). This suggests that 1) the accretion activity in high-redshift sources is unambiguously revealed thanks to the presence of a strong X--ray emission, and 2) the star-formation region as defined by Pope et al. (2008) contains not only objects in which the bolometric luminosity is dominated by the star-formation processes, but also a not-negligible number of objects hosting candidate obscured/Compton Thick AGN. In conclusion, our obscured AGN show Spitzer colors consistent with both an AGN dominated continuum and a starburst dominated continuum in the MIR. \\\\ \\par\\noindent We found that the hosts of obscured AGN are redder in (U-V) rest frame than the overall galaxy population at the same redshift: in particular, obscured AGN mainly populate the red sequence and the green valley in the color-magnitude plots (Figure 7, left panel), in agreement with the results of Nandra et al. (2007), and Silverman et al. (2007). For the MUSIC sample the U-V galaxy colors are strongly correlated with the K band absolute magnitude (Figure 8, left panel), and therefore with the galaxy stellar mass, with the most massive systems having a redder color. The hosts of the obscured AGN are therefore found in the red-massive tail of the distribution of optically selected galaxies in all three redshift bins considered. AGN feedback is often invoked as one of the main responsible for the observed galaxy colors (Nandra et al. 2007, Hasinger 2008). However, it is well known that the main ingredient for nuclear activity is the presence of a SMBH in galaxy nuclei, and that SMBHs are found nearly exclusively in massive galaxies (e.g. Magorrian et al. 1998). Therefore, it is not truly surprising to find AGN hosted in massive galaxies and the simple presence of AGN in massive red galaxies is not enough to argue for a significant feedback effect on the observed colors, because of the strong color-mass correlation. Were AGN feedback responsible for the observed red colors, since galaxy colors are strongly correlated with the galaxy mass and AGN are found preferably in massive galaxies, then AGN feedback should be also considered as one of the main players in the building of the galaxy mass-color correlation. \\\\ \\par\\noindent We found that about 2/3 of the obscured AGN hosts at all redshifts show substantial ($>10$ M$_\\odot$ yr$^{-1}$) star formation activity (Figure 9 left panel) and about half live in galaxies which are still actively forming stars with respect to their mass (Figure 9, right panel). For these sources, the observed red colors are likely due to dust extinction rather than evolved stellar population. We then conclude that a significant fraction of obscured AGN live in massive, dusty star-forming galaxies with red optical colors. This result is in qualitative agreement with the morphological analysis. Higher luminosity X-ray selected AGN are not systematically found in objects with the highest SSFR (see Fig. 9 right panel), in agreement with Alonso-Herrero et al. (2008). \\\\ \\par\\noindent We compared the number of obscured AGN and of all X--ray selected AGN to the number of field galaxies in broad bins of galaxy stellar mass (M$_*=10^{10}-10^{12}$ M$_\\odot$) and redshifts (z=0.6-1, z=1-2, z=2-4). We find that the AGN fraction increases with the host galaxy stellar mass, from $\\sim$1\\% at M$_*\\sim 10^{10}$ M$_\\odot$ to $\\sim$30\\% at M$_*\\sim 3\\times10^{11}$ M$_\\odot$ (see also Yamada et al. 2009), and the actual trend of increasing AGN fraction as a function of the stellar mass is probably steeper given the uncompleteness of the MUSIC sample at M$_{*}<10^{11}$ M$\\odot$ (see Section 4.4). The uncertainties on the stellar mass estimate from the SED fitting have the effect of shifting at lower masses (of $\\sim 0.25$ dex) the datapoints, leaving the total fraction and the slope unchanged. \\\\ We compared this trend with that observed in the local Universe (Best et al. 2005) for AGN with luminosity above similar thresholds. While the observed trend is the same, in all the investigated redshift bins the AGN fraction is higher than that observed in the local Universe, and it could likely be even higher. In fact, we are comparing here AGN selected with two different methods: forbidden emission line luminosity (SDSS) and X-ray emission (GOODS). The latter sample does not contain most Compton thick AGN. On the other hand, Compton thick AGN may well be present in [OIII] selected AGN samples. The fraction of Compton thick AGN not directly detected in deep Chandra surveys is estimated between 40\\% and 100\\% of the X-ray selected AGN, using infrared selection or other techniques (see Donley et al. 2008, Fiore et al. 2009 and references therein). Therefore, under the simplest assumption that this Compton Thick AGN fraction is constant with the galaxy mass, the discrepancy observed in Figure 10 can increase by up of a factor of 2. \\\\ The fraction of active galaxies to the total galaxy population is proportional to the AGN duty cycle. Our results would thus suggest higher AGN duty cycles at z=2-4 than at z=0, in agreement with expectations from most recent semi-analytic models (e.g. Menci et al. 2008), in which at higher redshift the AGN activity is present in a larger number of galaxies than locally. \\\\ \\par\\noindent The fact that the most luminous obscured AGN are found in the most massive galaxies at all investigated redshifts may suggest that the L/L$_{Edd}$ of the obscured AGN is similar, particularly in the case of the most luminous sources (logLx$>43$ erg s$^{-1}$), for which the threshold in luminosity introduces a bias against the sources accreting at lower rates in the lowest redshift bin. Assuming the local Magorrian relation between M$_{BH}$ and M$_*$ (e.g Marconi \\& Hunt 2003) and a bolometric correction of 20 (e.g. Marconi et al. 2004) the median values of L$_X$/M$_*$=32.83 and L$_X$/M$_*$=32.87 correspond to L/L$_{Edd}\\sim0.1$. Although suffering from large uncertainties associated with the stellar mass and BH mass estimates (see Section 4.4), this value can be considered as representative of the accretion state of the most luminous, obscured AGN in the present sample. Similar results are obtained from a comprehensive analysis of host galaxies properties of Chandra Deep Field North X--ray sources sources at z=2-4 (Yamada et al. 2009) and are also typical of unobscured Type 1 AGN at z$>1$ (Merloni et al. 2010, Trump et al. 2009). \\\\ It is instructive to compare these findings with similar estimates in the local Universe. Kauffmann \\& Heckman (2009) using SDSS data found an average L/L$_{Edd}$ value of $\\sim 0.01$ and a log normal distribution for this parameter, for AGN hosted in galaxies with significant on-going star formation, while AGN in inactive host galaxies follow a power-law distribution. As discussed in section 4.4, and shown in the upper right panel of figure 11, we also find marginal evidence of different accretion rates distributions for the populations of AGN in 'active' and 'inactive' host galaxies. \\\\" }, "0910/0910.3833_arXiv.txt": { "abstract": "In order to investigate whether galaxy structures are compatible with the predictions of the standard LCDM cosmology, we focus here on the analysis of several simple and basic statistical properties of the galaxy density field. Namely, we test whether, on large enough scales (i.e., $r>10$ Mpc/h), this is self-averaging, uniform and characterized by a Gaussian probability density function of fluctuations. These are three different and clear predictions of the LCDM cosmology which are fulfilled in mock galaxy catalogs generated from cosmological N-body simulations representing this model. We consider some simple statistical measurements able to tests these properties in a finite sample. We discuss that the analysis of several samples of the Two Degree Field Galaxy Redshift Survey and of the Sloan Digital Sky Survey show that galaxy structures are non self-averaging and inhomogeneous on scales of $\\sim$ 100 Mpc/h, and are thus intrinsically different from LCDM model predictions. Correspondingly the probability density function of fluctuations shows a \"fat tail\" and it is thus different from the Gaussian prediction. Finally we discuss other recent observations which are odds with LCDM predictions and which are, at least theoretically, compatible with the highly inhomogeneous nature of galaxy distribution. We point out that inhomogeneous structures can be fully compatible with statistical isotropy and homogeneity, and thus with a relaxed version of the Cosmological Principle. ", "introduction": "In the past twenty years many observations have been dedicated to the study of the large scale distributions of galaxies \\cite{cfa1,cfa2,pp,ssrs2,lcrs,sdss,colless01}. In particular during the last decade two ambitious observational programs have measured the redshift of more than one million objects \\cite{sdss,colless01}. All these surveys have detected larger and larger structures, thus finding that galaxies are organized in a complex network of clusters, super-clusters, filaments and voids. For instance the famous ``slice of the Universe'', that represented the first set of observations done for the CfA Redshift Survey in 1985 \\cite{dlhg85}, mapped spectroscopic observations of about 1100 galaxies in a strip on the sky 6 degrees wide and about 130 degrees long. This initial map was quite surprising, showing that the distribution of galaxies in space was anything but random, with galaxies actually appearing to be distributed on surfaces, almost bubble like, surrounding large empty regions, or ``voids.''. The structure running all the way across the survey between 50 and 100 Mpc/h\\footnote{We use $H_0=100 h$ km/sec/Mpc for the value of the Hubble constant.} was called the ``Great Wall'' and at the time of the discovery was the largest single structure detected in any redshift survey. Its dimensions, limited only by the sample size, are about $200\\times 80 \\times 10 $ Mpc/h, a sort of like a giant quilt of galaxies across the sky \\cite{gh89}. More and more galaxy large scale structures were identified in the other redshift surveys such as the Perseus-Pisces super-cluster \\cite{pp} which is one of two dominant concentrations of galaxies in the nearby universe. This long chain of galaxies lies next to the the so-called Taurus void, which is a large circular void bounded by walls of galaxies on either side of it. The void has a diameter of about 30 Mpc/h. Few years ago, in the larger sample provided by the Sloan Digital Sky Survey (SDSS), it has been discovered the Sloan Great Wall \\cite{sloangreatwall}, which is a giant wall of galaxies which may be the largest known structure in the Universe, being nearly three times longer than the Great Wall. The discovery of larger and larger structures was surprising because standard cosmological models unambiguously predict that fluctuations should be small on large scales with rapidly decaying correlations; i.e., the spatial extension of structures in these models should be limited to some tens Mpc/h. ", "conclusions": "We interpret the systematic differences found in the behavior of the PDF of conditional fluctuations as due to a systematic effect in the fact that sample volumes are not large enough for conditional fluctuations, filtered at such large scales, to be self-averaging, i.e. to contain enough structures and voids of large size to allow a reliable determination of average (conditional) quantities. We pointed out the problems related to the estimation of amplitude of fluctuations and correlation properties from statistical quantities which employ the normalization to the estimation of the sample average. As long as a distribution inside the given sample is not self-averaging, and thus not homogeneous, the estimation of the two-point correlation function is necessarily biased by strong finite size effects. Our results are incompatible with homogeneity at scales smaller than $\\sim 100$ Mpc/h. While these results are at odds with LCDM predictions, they can be compatible, at least theoretically, with a several recent observations which also pose fundamental challenges to such a model. For instance, Kashlinsky et al. \\cite{Kashlinsky}, by studying the fluctuations in the cosmic microwave background generated by the scattering of the microwave photons by the hot X-ray emitting gas inside clusters, have measured a coherent flow out to 300 Mpc/h with a fairly high amplitude of 600-$10^3$ km/sec. This is incompatible with the standard LCDM model predictions. Indeed, on such large scales the theoretical predictions are very simple, because in these models gravitational clustering is still linear at those scales as density fluctuations are small on large scales. Similarly Watkins et al. \\cite{Watkins} estimated the bulk flow in all major peculiar velocity surveys finding that the data suggest that the bulk flow within a Gaussian window of radius 50 Mpc/h is 407 km/s. They noticed that this large-scale bulk motion indicates that there are significant density fluctuations on very large scales. Indeed, a flow of this amplitude on such a large scale is not expected in the LCDM model cosmology, for which the predicted one-dimensional r.m.s. velocity is about 110 km/s. Thus the same problem for the model predictions we found in the galaxy density field, are found also for the galaxy velocity field. They may have the same origin, namely the fact that there are galaxy structures which are too extended in space and have a too large amplitude to be compatible with the predictions of the standard model. In the case of the velocity field there is another important element: the velocity field is generated by all the mass and not only by the luminous component. Thus if the velocity field is so high on such large scales, this may imply that this is generated by the large scale inhomogeneities present in the overall mass distribution, i.e. luminous plus dark. The fact that the whole mass distribution is not homogeneous is compatible also with the results on the matter density field derived by gravitational lens observations, where very extended structures have been found \\cite{massey}. Thus an important point which we aim to investigate in future works, concerns the characterizing of the gravitational field generated by galaxies. When a distribution is inhomogeneous important contributions to the gravitational force acting on a point can be due to faraway sources \\cite{force}. The relation between the large scale inhomogeneities of galaxy distribution and other observational data should be examined in detail. For instance a statistically significant anisotropy of the Hubble diagram at redshifts $z < 0.2$ was discovered by \\cite{sw08}. A local violation of statistical isotropy and homogeneity, which may very well happen when matter distribution is inhomogeneous \\cite{sl94,pwa}, can be related to such findings. although it is not excluded that a systematic error in the observations or data analysis affect these results. Finally, it is worth mentioning that Joyce et al. \\cite{pwa} pointed out that an inhomogeneous distribution, as a fractal, does not preclude the description of its gravitational dynamics in the framework of the Friedmann-Robertson-Walker solutions to general relativity. Indeed, the problem is often stated as being due to the incompatibility of a fractal with the Cosmological Principle, where this principle is identified with the requirement that the matter distribution be isotropic and homogeneous. This identification is in fact very misleading for a non-analytic and inhomogeneous structure like a fractal, in which all points are equivalent statistically, satisfying what has been called a Conditional Cosmological Principle \\cite{sl94,book}. By treating the fractal as a perturbation to an open cosmology in which the leading homogeneous component is the cosmic background radiation one may get a simple explanation for the supernovae data which indicate the absence of deceleration in the expansion. This is indeed a very simplified theoretical model to interpret the SN data and the large scale inhomogeneities in framework of the Friedmann-Robertson-Walker metric. \\begin{theacknowledgments} I am in debt with Yuri V. Baryshev and Nickolay L. Vasilyev for many collaborations on this topic. I also thank Tibor Antal, Michael Joyce, Andrea Gabrielli, Martin Lopez-Correidoira and Luciano Pietronero for useful remarks and discussions. I am grateful to Michael Blanton and David Hogg and for interesting comments. \\end{theacknowledgments}" }, "0910/0910.1836_arXiv.txt": { "abstract": "We present optical spectroscopic follow-up of a sample of Distant Red Galaxies (DRGs) with $K^{tot}_{s,Vega}<22.5$, selected by $(J-K)_{Vega}>2.3$, in the Hubble Deep Field South (HDFS), the \\1054 field, and the Chandra Deep Field South (CDFS). Spectroscopic redshifts were obtained for 15 DRGs. Only 2 out of 15 DRGs are located at $z<2$, suggesting a high efficiency to select high-redshift sources. From other spectroscopic surveys in the CDFS targeting intermediate to high redshift populations selected with different criteria, we find spectroscopic redshifts for a further 30 DRGs. We use the sample of spectroscopically confirmed DRGs to establish the high quality (scatter in $\\Delta z/(1+z)$ of $\\sim 0.05$) of their photometric redshifts in the considered deep fields, as derived with EAZY (Brammer et al. 2008). Combining the spectroscopic and photometric redshifts, we find that 74\\% of DRGs with $K^{tot}_{s,Vega} < 22.5$ lie at $z>2$. The combined spectroscopic and photometric sample is used to analyze the distinct intrinsic and observed properties of DRGs at $z<2$ and $z>2$. In our photometric sample to $K^{tot}_{s,Vega} < 22.5$, low-redshift DRGs are brighter in $K_s$ than high-redshift DRGs by 0.7 mag, and more extincted by 1.2 mag in $A_V$. Our analysis shows that the DRG criterion selects galaxies with different properties at different redshifts. Such biases can be largely avoided by selecting galaxies based on their rest-frame properties, which requires very good multi-band photometry and high quality photometric redshifts. ", "introduction": " ", "conclusions": "" }, "0910/0910.0001_arXiv.txt": { "abstract": "We use the ultra-deep WFC3/IR data over the HUDF and the Early Release Science WFC3/IR data over the CDF-South GOODS field to quantify the broadband spectral properties of candidate star-forming galaxies at $z\\sim7$. We determine the $UV$-continuum slope $\\beta$ in these galaxies, and compare the slopes with galaxies at later times to measure the evolution in $\\beta$. For luminous $L_{z=3}^{*}$ galaxies, we measure a mean $UV$-continuum slope $\\beta$ of $-2.0\\pm0.2$, which is comparable to the $\\beta\\sim-2$ derived at similar luminosities at $z\\sim5-6$. However, for the lower luminosity $0.1L_{z=3}^{*}$ galaxies, we measure a mean $\\beta$ of $-3.0\\pm0.2$. This is substantially bluer than is found for similar luminosity galaxies at $z$$\\sim$4, just 800 Myr later, and even at $z$$\\sim$5-6. In principle, the observed $\\beta$ of $-3.0$ can be matched by a very young, dust-free stellar population, but when nebular emission is included the expected $\\beta$ becomes $\\geq-2.7$. To produce these very blue $\\beta$'s (i.e., $\\beta\\sim-$3), extremely low metallicities and mechanisms to reduce the red nebular emission seem to be required. For example, a large escape fraction (i.e., $f_{esc}\\gtrsim0.3$) could minimize the contribution from this red nebular emission. If this is correct and the escape fraction in faint $z\\sim7$ galaxies is $\\gtrsim$0.3, it may help to explain how galaxies reionize the universe. ", "introduction": "The spectral properties of high-redshift galaxies must undergo dramatic changes at some point in the past, as the metallicities in these systems drop to lower values and these systems become progressively younger. In the limit of low metallicities, gas is no longer able to cool efficiently, likely resulting in massive extremely low-metallicity (or Population III) stars whose hot atmospheres are expected to result in a very hard $UV$-continuum spectrum and strong HeII 1640 emission. However, the strong UV flux coming from hot stars is expected to be largely offset by the redder nebular continuum light produced by the ionized gas surrounding these massive stars (e.g., Schaerer 2002). The newly installed WFC3/IR camera on the Hubble Space Telescope permits us to observe faint $z\\gtrsim7$ galaxies $\\gtrsim40\\times$ more efficiently than before, providing us with our most detailed look yet at the $UV$ light and spectral properties of $z\\gtrsim7$ galaxies. Already $\\gtrsim$25 likely $z$$\\sim$7-8 galaxies have been identified in the early WFC3/IR data over the Hubble Ultra Deep Field (HUDF: Oesch et al.\\ 2009b; Bouwens et al.\\ 2009b; McLure et al.\\ 2009; Bunker et al.\\ 2009), and $\\gtrsim$20 $z\\sim7$ candidate galaxies in the WFC3/IR Early Release Science (ERS) observations over the CDF-South (R.J. Bouwens et al.\\ 2009, in prep). Here we take advantage of these early WFC3/IR observations to study the spectral properties of candidate $z\\sim7$ galaxies. Our principal focus will be on the slope of $z\\sim7$ galaxy spectra in the $UV$-continuum -- since this slope is the primary observable we can derive from the available broadband imaging data with WFC3. The $UV$-continuum slope $\\beta$ ($f_{\\lambda}\\propto \\lambda^{\\beta}$: e.g., Meurer et al.\\ 1999) provides us with a powerful constraint on the age, metallicity, and dust content of high-redshift galaxies; it has already been the subject of much study at $z\\sim3-6$ (Lehnert \\& Bremer 2003; Stanway et al.\\ 2005; Yan et al.\\ 2005; Bouwens et al.\\ 2006; Hathi et al.\\ 2008; see Bouwens et al. 2009a for a systematic study at $z\\sim2-6$) and even at $z\\sim7$ (Gonzalez et al.\\ 2009) using NICMOS data. Throughout this work, we quote results in terms of the luminosity $L_{z=3}^{*}$ Steidel et al.\\ (1999) derived at $z\\sim3$: $M_{1700,AB}=-21.07$. We refer to the HST F606W, F775W, F850LP, F105W, F125W, and F160W bands as $V_{606}$, $i_{775}$, $z_{850}$, $Y_{105}$, $J_{125}$, and $H_{160}$, respectively, for simplicity. Where necessary, we assume $\\Omega_0 = 0.3$, $\\Omega_{\\Lambda} = 0.7$, $H_0 = 70\\,\\textrm{km/s/Mpc}$. All magnitudes are in the AB system (Oke \\& Gunn 1983). ", "conclusions": "Before discussing the extraordinarily blue $UV$-continuum slopes $\\beta\\sim-3$ found for lower luminosity galaxies at $z\\sim7$, we first consider the relatively luminous $L_{z=3}^{*}$ galaxy candidates. These sources have measured $\\beta$'s of $-2$, which is similar to that observed for luminous galaxies at $z\\sim3-6$ (Figure~\\ref{fig:mcolor}). Such slopes can be fit by a moderately young, subsolar ($0.2\\,Z_{\\odot}$) stellar population, with a maximum $E(B-V)$ of $\\sim$0.05 (Calzetti et al.\\ 2000 extinction law: see also Bouwens et al.\\ 2009a). The lower luminosity ($0.1L_{z=3}^{*}$) galaxy candidates at $z\\sim7$, by contrast, have observed $\\beta$'s of $-3.0\\pm0.2$. This is much bluer than for luminous $z\\sim7$ galaxies (by $\\Delta \\beta\\sim1$) and also much bluer than is found at $z\\sim5-6$ (Figures~\\ref{fig:colmag}-\\ref{fig:mcolor}). This makes these galaxies of great interest since they are likely to be even younger, more metal poor, and dust-free than any galaxies known. This is not a selection effect (\\S3.4), and cannot be attributed to Ly$\\alpha$ emission contributing to the broadband fluxes. Ly$\\alpha$ does not move into the $J_{125}$-band (used to estimate $\\beta$) until $z\\gtrsim8.1$, and $\\lesssim10$\\% of our $z$-dropout sample extends to $z>8$. An AGN contribution seem similarly unlikely, given the rarety of AGN signatures in faint Lyman-Break and Ly$\\alpha$-emitter samples (e.g., Nandra et al.\\ 2002; Ouchi et al.\\ 2008). What then is the explanation for the very blue $\\beta$'s? We explore several possibilities: \\begin{figure} \\epsscale{1.16} \\plotone{f3.eps} \\caption{$UV$-continuum slope $\\beta$ we would expect for $z\\sim7$ galaxies as a function of age for constant star formation (CSF) models and an instantaneous burst (Schaerer 2002). The gray band denotes the observed mean $\\beta$ and its uncertainty. The slopes $\\beta$ derived from the stellar light (dotted lines) and the stellar + nebular light (solid lines) are shown. While it is in principle possible to obtain $\\beta$'s of $-3$ with standard low metallicity ($\\geq$0.02 $Z_{\\odot}$) models, including the nebular emission associated with hot stars make the predicted $\\beta$'s $\\geq-2.7$ and hence too red.\\label{fig:dbeta}} \\end{figure} \\begin{figure} \\epsscale{1.16} \\plotone{f4.eps} \\caption{$UV$-continuum slope $\\beta$ Schaerer (2003) calculated as a function of age for instantaneous burst models for different metallicities. Both the slopes $\\beta$ derived from the stellar light (dotted lines) and the stellar + nebular light (solid lines) are shown. The pure stellar light (dotted lines) has very blue $UV$-continuum slopes $\\beta$'s ($\\beta\\lesssim-3$) for all the low metallicity cases considered here. Changing the IMF does not appear to change the conclusion here in any significant way. Of course, these same hot low metallicity stars also ionize the gas around them, thus producing a substantial amount of redder nebular continuum light. This makes the total SED of a galaxy much redder in general, and in the calculations by Schaerer (2003) shown here, $\\beta$ never becomes bluer than $-$3.0. \\label{fig:dbeta2}} \\end{figure} \\subsection{Standard stellar population models} A first question is whether it is possible to obtain a $\\beta$ of $\\sim-3$ using standard stellar population models (e.g., Leitherer et al.\\ 1999; Bruzual \\& Charlot 2003). The answer is that it is possible, but only for very young ($<$5 Myr) star-forming systems (see Figure~\\ref{fig:dbeta}). However, to do so ignores the nebular continuum emission from the ionized gas around the young stars. Including this nebular continuum emission can redden the observed $\\beta$ by as much as $\\Delta\\beta\\sim$0.5. Figure~\\ref{fig:dbeta} also shows the $\\beta$'s predicted for several low metallicity (0.02 $Z_{\\odot}$ and 0.2 $Z_{\\odot}$) starbursts, as a function of age for the Schaerer (2002) stellar population models (which -- like \\textit{Starburst99} Leitherer et al.\\ 1999 -- include a nebular contribution). In the best cases, the models predict $\\beta$'s as blue as $-2.7$, which is redder than what we observe in our faint samples.\\footnote{To match the wavelength baseline for $\\beta$ measured here, we added $-$0.1 to the $\\beta$'s (1300-1800\\AA baseline) tabulated in the Schaerer (2002, 2003).} This would suggest that lower metallicities are needed since the standard Leitherer et al.\\ (1999) or Bruzual \\& Charlot (2003) stellar population models do not produce blue enough colors to match those found in our lower luminosity $z\\sim7$ sample. Of course, the significance of this result is only modest ($<1.3\\sigma$), but the similarly blue colors observed for z$\\sim$8 selections (Section 3.6) suggest that we may want to take this finding at face value. \\subsection{Extremely low metallicities ($\\leq10^{-3}\\,\\,Z_{\\odot}$)} In Figure~\\ref{fig:dbeta2} we present the $\\beta$'s predicted by the Schaerer (2003) stellar population models (which conveniently provide predictions at extremely low metallicities) as a function of age for instantaneous bursts assuming a metallicity of $0.5\\times10^{-3} Z_{\\odot}$, $0.5\\times10^{-5} Z_{\\odot}$, and zero (population III).\\footnote{The use of instantaneous burst models allows us to explore the most extreme cases. Other models ($\\tau$, inverse $\\tau$, CSF) produce comparable $\\beta$'s.} From the figure, it is clear that the nebular component contributes significantly to the total light output from $\\lesssim10^{-3}$ $Z_{\\odot}$ stellar populations. While initially somewhat red due to the nebular contribution, the predicted $\\beta$'s for these ultra-low metallicity models become much bluer at ages $>10$ Myr, eventually reaching $\\beta$'s of $-$3. Such metallicities and ages are not necessarily unreasonable for lower luminosity galaxies at $z\\sim7$, and therefore at least one possible explanation for the very blue $\\beta$'s in our selections is that the metallicities for galaxies in our sample may be $\\lesssim10^{-3}\\,Z_{\\odot}$. The above explanation may explain some of the very blue galaxies in our selection, but given that $\\beta\\sim-3$ only for a limited period (10-30 Myr after a burst), it seems unlikely to be the general explanation (unless updated models revise the theoretical SEDs). \\subsection{Top-heavy IMF} One seemingly attractive explanation for the very blue $\\beta$'s observed is through a top-heavy IMF, since galaxy stellar populations would be weighted towards massive, blue stars. The difficulty with this explanation is that these same massive stars are extraordinarily efficient at ionizing the gas around them -- resulting in substantial nebular emission and leaving the galaxy with a net $\\beta$ no bluer (and likely redder) than the young ($<$1 Myr) bursts shown in Figures~\\ref{fig:dbeta}-\\ref{fig:dbeta2} (see also discussion in Leitherer \\& Heckman 1995). \\subsection{Minimizing nebular emission} A significant obstacle to matching the very blue $\\beta$'s observed is the red nebular emission associated with hot, ionizing stars. The nebular contribution could be reduced in a number of ways, by changes to the ionization parameter, metallicity, geometry, etc. Assessing the impact of such changes would benefit from further detailed modelling. However, we should emphasize that regardless of changes to the nebular contribution very low metallicity models (or very young ages) appear to be needed to match the very blue $\\beta$'s observed. \\begin{figure} \\epsscale{1.16} \\plotone{f5.eps} \\caption{Mean $UV$-continuum slopes $\\beta$'s predicted for high-redshift galaxy samples versus the escape fraction $f_{esc}$. The predictions are made using the Schaerer (2003) stellar population models (including nebular emission) for metallicities of $<10^{-3}$ $Z_{\\odot}$. Galaxies are assumed to be observed at some random point during their star formation histories and to experience instantaneous bursts of star formation every 10-40 Myr (the dashed, solid, and dotted lines give the average $\\beta$ for the first 10, 20, and 40 Myr, respectively, of the instantaneous burst). For $f_{esc}=1$, all of the ionizing radiation from a galaxy escapes into the IGM and hence does not contribute to ionizing the gas within a galaxy (and hence the contribution from nebular continuum emission to the total light is minimal). On the other hand, for $f_{esc}\\sim0$ (preferred in some simulations: e.g., Gnedin et al.\\ 2008), the ionizing radiation from the hot stars does not make it out of galaxies -- resulting in substantial nebular continuum emission (see Figure~\\ref{fig:dbeta2}) and hence a much redder $\\beta$. Perhaps the very blue $\\beta$'s observed (\\textit{shaded gray region}) could indicate that the escape fraction is larger at $f_{esc}\\gtrsim0.3$ than what has commonly been considered? \\label{fig:escapef}} \\end{figure} \\subsection{Changes in Escape Fraction?} One possibility that we have explored to reduce the nebular emission contribution is if the ionizing radiation leaks directly into the IGM (e.g., due to the effect of SNe on the galaxies' ISM, possibly from a top-heavy IMF: Trenti \\& Shull 2009). We consider such a possibility schematically in Figure~\\ref{fig:escapef}, showing the time-averaged $\\beta$'s expected for stellar populations of various metallicities as a function of the escape fraction $f_{esc}$ of ionizing photons into the IGM. For $f_{esc}$ of unity, we simply recover the $\\beta$'s from the pure stellar SEDs and for $f_{esc}$ of zero, we recover the stellar + nebular SEDs. For fractional $f_{esc}$, we interpolate between the two extremes. The time-averaged $\\beta$'s used for Figure~\\ref{fig:escapef} is based upon those presented in Figure~\\ref{fig:dbeta2}. Comparing the predicted $\\beta$'s with that observed (grey-shaded region: Figure~\\ref{fig:escapef}), we see that $f_{esc}\\gtrsim$0.3 would permit us to easily match the observations. Such an escape fraction is signi\ufb01cantly higher than the $\\sim$10\\% frequently assumed in calculations assessing the sufficiency of galaxies to reionize the universe. Since most of the luminosity density at $z>7$ comes from low luminosity galaxies, the estimated number of ionizing photons in the $z>7$ universe could increase by factors of $\\gtrsim$3. Such a large change could provide the needed photons to reionize the universe, providing a resolution to the current debate (e.g., Bouwens et al. 2008; Oesch et al. 2009a, 2009b; McLure et al. 2009; Bunker et al. 2009; Gonzalez et al. 2009; Ouchi et al.\\ 2009; Pawlik et al.\\ 2009)." }, "0910/0910.2232_arXiv.txt": { "abstract": "We study the evolution of black holes (BHs) on the $M_{\\rm BH}-\\sigma$ and $M_{{\\rm BH}}-M_{\\rm bulge}$ planes as a function of time in disk galaxies undergoing mergers. We begin the simulations with the progenitor black hole masses being initially below $(\\Delta \\log M_{\\rm BH,i}\\sim -2)$, on $(\\Delta \\log M_{\\rm BH,i}\\sim 0)$ and above $(\\Delta \\log M_{\\rm BH,i}\\sim 0.5)$ the observed local relations. The final relations are rapidly established after the final coalescence of the galaxies and their BHs. Progenitors with low initial gas fractions ($f_{\\rm gas}=0.2$) starting below the relations evolve onto the relations $(\\Delta \\log M_{\\rm BH,f}\\sim -0.18)$, progenitors on the relations stay there $(\\Delta \\log M_{\\rm BH,f}\\sim 0)$ and finally progenitors above the relations evolve towards the relations, but still remaining above them $(\\Delta \\log M_{\\rm BH,f}\\sim 0.35)$. Mergers in which the progenitors have high initial gas fractions ($f_{\\rm gas}=0.8$) evolve above the relations in all cases $(\\Delta \\log M_{\\rm BH,f}\\sim 0.5)$. We find that the initial gas fraction is the prime source of scatter in the observed relations, dominating over the scatter arising from the evolutionary stage of the merger remnants. The fact that BHs starting above the relations do not evolve onto the relations, indicates that our simulations rule out the scenario in which overmassive BHs evolve onto the relations through gas-rich mergers. By implication our simulations thus disfavor the picture in which supermassive BHs develop significantly before their parent bulges. ", "introduction": "Observations in recent years have revealed a strong correlation in the local Universe between the central supermassive black holes (BHs) and their host galaxies as manifested in the relation between the BH mass and the bulge velocity dispersion, $M_{\\rm BH}-\\sigma$ (e.g. \\citealp{2000ApJ...539L...9F,2002ApJ...574..740T,2009ApJ...698..198G};), the bulge stellar mass $M_{{\\rm BH}}-M_{\\rm bulge}$ (e.g. \\citealp{1998AJ....115.2285M}; \\citealp{2004ApJ...604L..89H}), the concentration of light in the galaxy (e.g. \\citealp{2001ApJ...563L..11G}) and the bulge binding energy, $M_{{\\rm BH}}-M_{\\rm bulge}\\sigma^{2}$ (e.g. \\citealp{2007ApJ...665..120A}). The evolution of these relations with redshift is still unclear with some studies finding evolution in the $M_{\\rm BH}-\\sigma$ \\citep{2006ApJ...645..900W,2008ApJ...681..925W} and $M_{{\\rm BH}}-M_{\\rm bulge}$ (\\citealp{2006ApJ...649..616P,2007ApJ...667..117T,2009arXiv0911.2988D}) relations, with the high redshift BHs being overmassive for a fixed $\\sigma$ and $M_{\\rm bulge}$ compared to the local relations, whereas other studies are consistent with no redshift evolution in the observed correlations \\citep{2009arXiv0908.0328G,2009ApJ...706L.215J}. A possible explanation for this discrepancy lies in the uncertainties in observational selection biases and in the evolution in the intrinsic scatter that is typically stronger for larger BH masses \\citep{2007ApJ...670..249L}. The observed correlations are typically explained using theoretical models relying on some form of self-regulated BH mass growth, in which gas is fed to the central black hole until the black hole releases sufficient energy to unbind the gas and blow it away in momentum- or pressure-driven winds (e.g. \\citealp{1998A&A...331L...1S,1999MNRAS.308L..39F,2001ApJ...554L.151B,2007ApJ...665.1038C,2009ApJ...699...89C}). The observed correlations and their evolution with redshift have been reproduced in semi-analytic models (e.g. \\citealp{2006MNRAS.365...11C,2008MNRAS.391..481S}), in self-consistent numerical simulations of both isolated galaxies and galaxy mergers (\\citealp{2005Natur.433..604D,2005MNRAS.361..776S,2006ApJ...641...90R}) as well as in galaxies simulated in a full cosmological setting (\\citealp{2007MNRAS.380..877S,2008ApJ...676...33D,2009MNRAS.398...53B}). The key assumption in these models is that the galaxies undergo a brief radiatively-efficient quasar phase triggered by gas-rich galaxy mergers (e.g. \\citealp{2002ApJ...580...73T,2006ApJS..163....1H,2008ApJS..175..356H}) during which the bulk of the BH growth is taking place and the observed correlations are established. In a previous paper (\\citealp{2009ApJ...690..802J}, hereafter J09), we showed that the merger remnants of both equal- and unequal-mass mergers of disk and elliptical galaxies satisfy the observed $M_{\\rm BH}-\\sigma$ and $M_{{\\rm BH}}-M_{\\rm bulge}$ correlations. In this Letter, we study for the first time in detail the evolution of the BH scaling relations during disk galaxy mergers using a new sample of high resolution simulations including a self-consistent formulation for BH feedback. We seed the BHs initially with masses corresponding to locations below, on and above the observed relations. Thus, in contrast to previous studies we study here for the first time also overmassive BHs lying initially above the observed relations. The BH scaling relations are defined for merger remnants that have reached their final dynamical state and it is not obvious if the scaling relations are valid during the merging process and at what stage the galaxies evolve onto the relations. \\begin{figure} \\centering \\includegraphics[width=7.4cm]{./Figures/f1.eps} \\caption{The relative distance between the BHs (top), the total SFR (middle), and the evolution of the total BH accretion rate (bottom) as a function of time. The star symbol indicate the time of BH coalescence and the filled circles at the bottom of the plots indicate the time at which ($M_{\\rm BH}$,$\\sigma$,$M_{\\rm bulge}$) are evaluated in Figs. \\ref{MBH_evo_new_BBH}-\\ref{MBH_evo_new_ABH}.} \\label{Dist_SFR_BH-evo} \\end{figure} \\begin{table*} \\caption{The simulated merger sample} \\scriptsize{ \\label{sims} \\centering \\begin{tabular}{c c c c c c c c c c c c c c} \\hline\\hline Model & $M_{\\rm BH,i1}$\\footnote[$a$]{Initial BH mass in $ 10^{5} M_{\\odot}$} & $M_{\\rm BH,i2}$\\footnotemark[$a$] & Mass ratio & $\\alpha$ & $f_{\\rm gas}$ & $N_{\\rm{gas,tot}}$ & $N_{\\rm{disk,tot}}$ & $N_{\\rm{bul,tot}}$ & $N_{\\rm{DM,tot}}$ & $M_{\\rm BH,f}$\\footnote[$b$]{Final BH mass in $ 10^{5} M_{\\odot}$} & $\\sigma_{\\rm bul,f}$\\footnote[$c$]{Final stellar velocity dispersion in $\\rm km/s$} & $M_{\\rm bul,f}$\\footnote[$d$]{Final bulge mass in $10^{10} M_{\\odot}$} \\\\ \\hline 11B2BH & 1.0 & 1.0 & 1:1 & 25 & 0.2 & 120 000 & 480 000 & 200 000 & 800 000 & 459 & 183.8 & 5.07 \\\\ 31B2BH & 1.0 & 1.0 & 3:1 & 25 & 0.2 & 80 000 & 320 000 & 133 333 & 533 333 & 271 & 151.8 & 1.90 \\\\ 31B8BH & 1.0 & 1.0 & 3:1 & 25 & 0.8 & 320 000 & 80 000 & 133 333 & 533 333 & 1136 & 181.6 & 2.97 \\\\ 31B2BH1 & 1.0 & 1.0 & 3:1 & 100 & 0.2 & 80 000 & 320 000 & 133 333 & 533 333 & 82.5 & 144.7 & 1.26 \\\\ \\hline 11O2BH & 159 & 159 & 1:1 & 25 & 0.2 & 120 000 & 480 000 & 200 000 & 800 000 & 600 & 174.6 & 3.38 \\\\ 31O2BH & 159 & 36.4 & 3:1 & 25 & 0.2 & 80 000 & 320 000 & 133 333 & 533 333 & 253 & 132.0 & 1.55 \\\\ 31O8BH & 159 & 36.4 & 3:1 & 25 & 0.8 & 320 000 & 80 000 & 133 333 & 533 333 & 1698 & 155.2 & 3.02 \\\\ \\hline 11A2BH & 477 & 477 & 1:1 & 25 & 0.2 & 120 000 & 480 000 & 200 000 & 800 000 & 1200 & 164.2 & 3.35 \\\\ 31A2BH & 477 & 109 & 3:1 & 25 & 0.2 & 80 000 & 320 000 & 133 333 & 533 333 & 652 & 128.7 & 1.59 \\\\ 31A8BH & 477 & 109 & 3:1 & 25 & 0.8 & 320 000 & 80 000 & 133 333 & 533 333 & 842 & 122.3 & 1.21 \\\\ \\hline \\end{tabular} } \\end{table*} \\begin{figure*} \\centering \\includegraphics[width=14.9cm]{./Figures/f2.eps} \\caption{The evolution of the BHs in the B-merger sample $(\\Delta \\log M_{\\rm BH,i}\\sim -2)$ as a function of time on the $M_{\\rm BH}-\\sigma$ plane (top left panels) overplotted by lines giving the observed correlation from \\citet{2002ApJ...574..740T}. The bottom left panel gives the evolution on the $M_{\\rm BH}-M_{\\rm bulge}$ plane overplotted by the observed correlation from \\citet{2004ApJ...604L..89H}. Each point is separated by $\\Delta t=0.2 \\ \\rm Gyr$, with filled circles indicating the primary galaxy, triangles indicating the secondary galaxy, stars showing the time of BH coalescence and filled squares the final state. The panels on the right give the logarithmic offset (Eq. \\ref{eq:offset}) from the corresponding observed relation as a function of time.} \\label{MBH_evo_new_BBH} \\end{figure*} ", "conclusions": "In this Letter, we have studied the evolution of galaxy mergers on the $M_{\\rm BH}-\\sigma$ and $M_{{\\rm BH}}-M_{\\rm bulge}$ planes as a function of time. We have shown that progenitors with low initial gas fractions ($f_{\\rm gas}=0.2$) starting below the relations evolve onto the relations, progenitors on the relations stay there and progenitors above the relations evolve towards the relations, but still remaining above them. Progenitors with high initial gas fractions ($f_{\\rm gas}=0.8$) evolve above the relations in all cases, with the initial gas fraction thus being the prime source of scatter in the observed relations (see also \\citealt{2007ApJ...669...45H}). The evolution for all mergers is initially slow with the observed relations typically being rapidly established during a relatively short phase centered at the time of final coalescence of the BHs. The observations of \\citet{2006ApJ...645..900W,2008ApJ...681..925W} indicate that the BHs at high redshifts were typically overmassive by a factor of $\\sim3$ for a fixed $\\sigma$ and $M_{\\rm bulge}$ compared to the local relation. These BHs could plausibly have been formed during a brief quasar phase (e.g. \\citealp{2006ApJS..163....1H}) triggered during the mergers of very gas-rich galaxies, thus resulting in overmassive BH masses for their given $\\sigma$ and $M_{\\rm bulge}$, as seen in our $f_{\\rm gas}=0.8$ simulation series. However, this being the case it is not obvious how these galaxies would evolve onto the local observed relations until the present day. Another binary merger with massive BHs in place does not bring the galaxies onto the relations (O- and A-series), with high gas fractions mergers moving them even further away from the relations. Potentially, the bulge mass could be increased by dry minor mergers (e.g. \\citealp{2009ApJ...699L.178N}), whereas internal secular processes could be responsible for increasing the velocity dispersion. However, it is not obvious how both the velocity dispersion and the bulge mass could be increased at a fixed BH mass. Finally, another possibility could be that some of the BHs undergoing mergers are ejected in a sling-shot effect \\citep{1974ApJ...190..253S}, thus decreasing the total BH mass. The fact that the BHs starting below the relations evolve onto the relation, whereas the ones above do not, indicates that our simulations rule out the scenario in which overmassive BHs evolve onto the relations through gas-rich mergers, whereas undermassive BHs can evolve onto the relations in galaxy mergers. Thus, given the numerical limitations inherit in our simulations, our results disfavor the picture in which supermassive BHs develop significantly before their parent bulges. Finally, our current BH accretion and feedback prescription seems to describe the growth of BH during the merger phase adequately. However, it is not obvious that the present description is also valid during the secularly driven low accretion phase after the merger is completed. Some initial steps to improve the prescription have been taken by developing an entirely new momentum driven feedback prescription \\citep{2009arXiv0909.2872D}. Recent observations \\citep{2007MNRAS.382.1415S,2009ApJ...692L..19S} have indicated that there is a time lag of $\\sim0.5 \\ \\rm Gyr$ between the peak of star formation and the onset of AGN activity. Our current model has difficulties in reproducing this and thus developing models that shed light on this discrepancy might also ultimately help us understanding in how, when and why the BHs in galaxies evolve onto the observed relations." }, "0910/0910.5624_arXiv.txt": { "abstract": "{} {By studying the photospheric abundances of 4 RV\\,Tauri stars in the LMC, we test whether the depletion pattern of refractory elements, seen in similar Galactic sources, is also common for extragalactic sources. Since this depletion process probably only occurs through interaction with a stable disc, we investigate the circumstellar environment of these sources. } {A detailed photospheric abundance study was performed using high-resolution UVES optical spectra. To study the circumstellar environment we use photometric data to construct the spectral energy distributions of the stars, and determine the geometry of the circumstellar environment, whereas low-resolution Spitzer-IRS infrared spectra are used to trace its mineralogy.} {Our results show that, also in the LMC, the photospheres of RV\\,Tauri stars are commonly affected by the depletion process, although it can differ significantly in strength from source to source. From our detailed disc modelling and mineralogy study, we find that this process, as in the Galaxy, appears closely related to the presence of a stable Keplerian disc. The newly studied extragalactic objects have similar observational characteristics as Galactic post-AGB binaries surrounded by a dusty disc, and are therefore also believed to be part of a binary system. One source shows a very small infrared excess, atypical for a disc source, but still has evidence for depletion. We speculate this could point to the presence of a very evolved disc, similar to debris discs seen around young stellar objects.} {} ", "introduction": "\\label{introduction} With a distance of $\\sim$\\,50\\,kpc \\citep{feast99}, the Large Magellanic Cloud (LMC) is one of the best galaxies to study stellar evolution. The known distance allows us to calculate the luminosity for stellar sources and it is the ideal laboratory to study physical and chemical processes in an environment with a sub-solar metallicity of $Z\\sim 0.3-0.5$\\,Z$_\\odot$ \\citep{westerlund97}. Moreover, it is close enough to allow detailed studies of individual sources using large-aperture ground based telescopes. Here we focus on a sample RV\\,Tauri stars in the LMC. RV\\,Tauri stars are pulsating evolved stars with a characteristic light curve showing alternating deep and shallow minima. They are located in the high luminosity end of the Population II Cepheid instability strip \\citep{wallerstein02}. The post-AGB status of RV\\,Tauri stars was a long standing debate, but the detection of circumstellar dust around many objects \\citep{jura86}, and the detection of extragalactic pulsators and their large derived absolute luminosities, were in line with the expected evolutionary tracks of post-AGB stars. The first extragalactic RV\\,Tauri stars in the LMC were discovered by the MACHO experiment \\citep{alcock98}. \\citet{reyniers07a} and \\citet{reyniers07b} performed a chemical study on two LMC RV\\,Tauri stars, selected from those reported by \\citet{alcock98}, and found that, like in the Galaxy \\citep{vanwinckel03}, post-AGB stars are chemically much more diverse than previously anticipated. One of the stars, MACHO\\,47.2496.8, proved to be strongly enhanced in s-process elements, in combination with a very high carbon abundance (C/O\\,$>$\\,2 and $^{12}$C/$^{13}$C\\,$\\approx$\\,200). The metallicity of [Fe/H]\\,$=$\\,$-1.4$ is surprisingly low for a field LMC star. The s-process enrichment is large: the light s-process elements of the Sr-peak are enhanced by 1.2 dex compared to iron ([ls/Fe]\\,$=$\\,$+1.2$), while for the heavy s-process (Ba-peak) elements, an even stronger enrichment is measured: [hs/Fe]\\,$=$\\,$+2.1$. Lead was not found to be strongly enhanced. The patterns can only be understood assuming a very low efficiency of the $^{13}$C pocket \\citep{bonacic07} which is created during the dredge-up phenomenon and the associated partial mixing of protons into the intershell. It was the first detailed study of the s-process of a post-AGB star in an external galaxy. Another object, MACHO\\,82.8405.15, turned out to be chemically altered by the depletion phenomenon \\citep{reyniers07a}. Depletion of refractory elements in the photosphere is a chemical process in which chemical elements with a high dust condensation temperature are systematically underabundant \\citep[e.g.][]{maas05,giridhar05}. The special photospheric abundance patterns are the result of gas-dust separation in the circumstellar environment, followed by re-accretion of only the gas, which is poor in refractory elements. The photospheres become deficient in refractories (as Fe, Ca and the s-process elements), while the non-refractories are not affected. The best abundance tracers of the depletion phenomenon are the Zn/Fe and S/Ti ratios because the elements involved in either ratio have a similar nucleosynthetic formation channel, but have very different condensation temperatures. Intrinsically Fe-poor objects have [Zn/Fe] and [S/Ti] close to solar, which is not the case for depleted objects. With [Fe/H]\\,$=$\\,$-$2.6, in combination with [Zn/Fe]\\,$=$\\,$+$2.3 and [S/Ti]\\,$=$\\,$+$2.5, in MACHO\\,82.8405.15, there is no doubt that the depletion affected the photosphere of this LMC star very strongly. The very low abundances, as well as the clear correlation with condensation temperature, are shown in Figure~\\ref{depletie}, which is a reproduction of the figure in \\citet{reyniers07a}. Photospheric depletion is surprisingly common in Galactic post-AGB stars \\citep[e.g.][and references therein]{giridhar05,maas05}. In almost all depleted post-AGB objects, there is observational evidence that a stable circumbinary disc is present \\citep{deruyter06, vanwinckel07}. The discs are very compact \\citep[e.g.][]{deroo06, deroo07c} and very likely only found around binary post-AGB stars \\citep[e.g.][and references therein]{vanwinckel09}. Also in their infrared spectra, the RV\\,Tauri stars show unique spectral features. One of the best studied Galactic RV\\,Tauri stars is AC\\,Her \\citep{molster99}. Infrared studies with ISO have shown very strong crystalline silicate bands from 10--50\\,$\\mu$m. Recent studies with Spitzer show that this strong crystallinity is commonly observed, also in other post-AGB binary sources \\citep{gielen08}. Thanks to the efficient infrared detectors of the Spitzer satellite, very sensitive infrared observations allow us to probe for circumstellar dust, even around individual objects in external galaxies. The SAGE (Surveying the Agents of a Galaxy's Evolution) Spitzer LMC survey \\citep{meixner06} mapped the LMC, using all photometric bands of the Spitzer IRAC (InfraRed Array Camera, \\citealp{fazio04}) and MIPS (Multiband Imaging Photometer for Spitzer, \\citealp{rieke04}) instruments. This survey resulted in the detection of over 4 million sources. Thanks to the release of this database, we found that the LMC RV\\,Tauri stars as discovered with the MACHO experiment in the visible, indeed have infrared excesses with very similar SED shapes as many Galactic post-AGB binaries (see Fig.~\\ref{kleurkleur}). Since only those LMC sources with strong enough optical fluxes were selected, some observational bias exists to stars where we see the disc more face-on, since edge-on disc would obscure the star too much. \\begin{figure} \\centering \\resizebox{10cm}{!}{\\includegraphics{12982fg1.ps}} \\caption{Colour-colour diagram indicating the Galactic post-AGB sources with discs (grey circles) and extragalactic RV\\,Tauri stars as presented by \\citet{alcock98} (black circles). The grey dots represent the LMC objects as found by the SAGE-LMC survey \\citep{meixner06}.} \\label{kleurkleur} \\end{figure} \\begin{table*} \\caption{List of stellar parameters for our sample sources.} \\label{stelpar} \\vspace{0.cm} \\hspace{-.5cm} \\begin{tabular}{lllcrrccccc} \\hline \\hline Name & other name & $\\alpha$ (J2000) & $\\delta$ (J2000) & $T_{\\rm eff}$ & $\\log g$ & [Fe/H] & $E(B-V)_{tot}$ & $L_{\\rm IR}/L_*$ & $L_*$ & $P$\\\\ & &(h m s) & ($^\\circ$ ' '') & (K) & (cgs) & & & (\\%) & L$_\\odot$ & (days)\\\\ \\hline MACHO\\,79.5501.13 & J051418.1-691235 & 05 14 18.1 & -69 12 34.9 & 5750 & 0.5 & -2.0 & $0.14\\pm0.01$ & $60\\pm3$ & $5000\\pm500$ & 48.5\\\\ & HV\\,915 &&&&&&&&&\\\\ MACHO\\,81.8520.15 & J053254.5-693513 & 05 32 54.5 & -69 35 13.2 & 6250 & 1.0 & -1.5 & $0.27\\pm0.03$ & $0\\pm1$ & $4200\\pm500$ & 42.1\\\\ MACHO\\,81.9728.14 & J054000.5-694214 & 05 40 00.5 & -69 42 14.6 & 5750 & 1.5 & -1.0 & $0.05\\pm0.02$ & $53\\pm3$ & $4200\\pm500$ & 47.1\\\\ MACHO\\,82.8405.15 & J053150.9-691146 & 05 31 51.0 & -69 11 46.4 & 6000 & 0.5 & -2.5 & $0.05\\pm0.01$ & $84\\pm3$ & $4000\\pm500$ & 46.5\\\\ \\hline \\end{tabular} \\begin{footnotesize} \\begin{flushleft} Note: The name, equatorial coordinates $\\alpha$ and $\\delta$ (J2000), effective temperature $T_{\\rm eff}$, surface gravity $\\log g$ and metallicity [Fe/H] of our sample stars. Also given are the total reddening $E(B-V)_{tot}$, the energy ratio $L_{\\rm IR}/L_*$, the calculated luminosity (computed by integrating the dereddened photosphere), assuming a distance of $d=50$\\,kpc, and the period $P$ as given in \\citep{alcock98}. \\end{flushleft} \\end{footnotesize} \\end{table*} \\begin{table} \\caption{\\label{logUves}Log of the UVES observations and the obtained final S/N at a given spectral band. } \\centering \\vspace{0.5cm} \\hspace{0.cm} \\begin{tabular}{lrrrr} \\hline\\hline\\rule[0mm]{0mm}{3mm} Star & Exp. Time & Wavelength & S/N \\\\ & (sec.) & (nm) & \\\\ \\hline MACHO\\,79.5501.13 & 3600 & 375.8 -- 498.3 & 70 \\\\ & & 670.5 -- 1008.4 & 100 \\\\ MACHO\\,81.8520.15 & 3600 & 478.0 -- 680.8 & 70 \\\\ MACHO\\,81.9728.14 & 3600 & 478.0 -- 680.8 & 65 \\\\ \\hline \\end{tabular} \\begin{footnotesize} \\begin{flushleft} Note: The S/N is measured in the middle of the spectral window covered. \\end{flushleft} \\end{footnotesize} \\end{table} In this contribution we focus on the MACHO RV\\,Tauri objects in the LMC as detected by the MACHO experiment \\citep{alcock98}, and we connect the chemical studies based on high-resolution optical spectroscopy to the SED energetics and the infrared spectra obtained by Spitzer. Prior to this study, only 1 depleted post-AGB star in the LMC was known. Here we analyse the abundances of 4 more similar sources, testing if depletion is also a common process in the RV\\,Tauri stars of the LMC. As the distance to the LMC is known, we are able to discuss the evolutionary status of these extragalactic sources using their accurate position in the H-R diagram, which sofar has been impossible for Galactic post-AGB stars. This research was possible thanks to the SAGE-Spec international program (\\textit{http://sage.stsci.edu/}). The goal of this large spectroscopic infrared program is to complement the wealth of photometric data from the SAGE photometric survey, with an extensive spectroscopic follow-up programme using the infrared IRS spectrograph aboard of Spitzer (Kemper et al., 2009, submitted). The main goal of the survey is to determine the composition, origin and evolution, and observational characteristics of interstellar and circumstellar dust and its role in the LMC. A total of about 200 stars, in all stages of stellar evolution, HII and diffuse regions were observed in Spitzer-IRS\tand MIPS-SED mode. ", "conclusions": "Clearly the formation of stable dusty discs and the significant feedback of this disc on the central star is not exclusive to our Galaxy alone. Four out of five LMC RV\\,Tauri objects, for which we have high-resolution data, are found to be strongly affected by the depletion process in which the atmospheres became poor in refractory elements. Moreover, the infrared colours and spectral data show that three sources are surrounded by a highly processed, stable disc in which dust processing has been efficient. In one source PAH particles are formed, while there is no other evidence for intrinsic carbon enhancement in the photosphere. This object is intrinsically the most metal poor object of the sample. For MACHO\\,81.8520.15 we only found a small dust excess at 8\\,$\\mu$m and we could interpret this as evidence for strong disc evolution. In this star, even the Zn abundance is affected by depletion which means that the gas-dust separation occurred at low temperatures. Overall the RV\\,Tauri stars in the LMC display many characteristics of their Galactic peers. Whether these extragalactic stars also reside in a binary system remains unclear with the available data at hand. The low magnitude of these stars does not allow for a long-term radial velocity monitoring programme but, given the strong similarities to the Galactic binaries, a binary evolutionary channel is very likely needed to understand the RV\\,Tauri stars in the LMC as well." }, "0910/0910.2468_arXiv.txt": { "abstract": "We present the results of an X-ray mass analysis of the early-type galaxy NGC~4636, using \\Chandra\\ data. We have compared the X-ray mass density profile with that derived from a dynamical analysis of the system's globular clusters (GCs). Given the observed interaction between the central active galactic nucleus and the X-ray emitting gas in NGC~4636, we would expect to see a discrepancy in the masses recovered by the two methods. Such a discrepancy exists within the central $\\sim$10\\,kpc, which we interpret as the result of non-thermal pressure support or a local inflow. However, over the radial range $\\sim$\\,10--30\\,kpc, the mass profiles agree within the 1\\,$\\sigma$ errors, indicating that even in this highly disturbed system, agreement can be sought at an acceptable level of significance over intermediate radii, with both methods also indicating the need for a dark matter halo. However, at radii larger than 30\\,kpc, the X-ray mass exceeds the dynamical mass, by a factor of 4--5 at the largest disagreement. A Fully Bayesian Significance Test finds no statistical reason to reject our assumption of velocity isotropy, and an analysis of X-ray mass profiles in different directions from the galaxy centre suggests that local disturbances at large radius are not the cause of the discrepancy. We instead attribute the discrepancy to the paucity of GC kinematics at large radius, coupled with not knowing the overall state of the gas at the radius where we are reaching the group regime ($>$30\\,\\kpc), or a combination of the two. ", "introduction": "\\label{sec:intro} The current paradigm of galaxy formation describes how galaxies form embedded in massive dark matter halos. Whereas the measurement of rotation curves can be successfully applied to late-type galaxies to infer the presence of this dark matter \\citep[see e.g.][for a review]{sofue01}, this cannot be employed in early-type galaxies as their stars and gas are not supported by rotation. Therefore, different methods must be invoked to measure the galaxy mass. It has long been known that early-type galaxies contain hot ($\\sim$10$^{6}$K) X-ray emitting gas \\citep{forman85}, the temperature and density of which allow the determination of the total gravitating mass, assuming hydrostatic equilibrium and spherical symmetry. This approach has proved successful at yielding meaningful mass profiles for early-type galaxies \\citep[e.g.][]{osullivan04b,fukazawa06,humphrey06,zhang07}. The effect of the assumption of spherical symmetry has been addressed in the case of galaxy clusters, indicating that although compression and elongation along the line-of-sight can under or over-estimate the central mass respectively, this is only a small effect at large radius \\citep{piffaretti03}. However, the validity of the intrinsic assumption of hydrostatic equilibrium has been questioned with specific reference to early-type galaxies \\citep{diehl06}. NGC~4636 presents an ideal test-bed in this respect, as it is a highly disturbed system, with evidence of bubbles and shocks caused by previous AGN outbursts \\citep{jones02,ohto03,osullivan05b}. The use of dynamical tracers of the gravitational potential is also a well-established method to recover the kinematics of bound systems, both on the scale of globular clusters \\citep[e.g.][in M15]{gebhardt00}, and for the Galaxy itself \\citep{chakrabarty01,genzel00,ghez98}. The use of globular clusters (GCs) as tracers of the potential has been particularly successful in recovering mass profiles of nearby elliptical galaxies \\citep[examples include][]{romanowsky01,cote03,bergond06,schuberth06,woodley07}. Similarly, dedicated surveys of planetary nebulae in early-type galaxies can also be used to derive the distribution of matter \\citep{douglas07}, although care is required to avoid complications from distinct populations of planetary nebulae, which have been seen for example in the galaxy NGC 4697 \\citep{sambhus06}. These approaches involve solving the Jeans equations under the assumption of spherical symmetry to determine the galaxy mass. Interestingly, a study of three early-type galaxies using planetary nebulae kinematics, \\citet{romanowsky03} concluded a significant lack of dark matter in these systems. However, \\citet{dekel05} showed these data to be consistent with a massive dark halo when more radial orbits were considered. This highlights the mass-anisotropy degeneracy present in this approach, which can be broken by considering higher order velocity moments \\citep{lokas07}. A further systematic effect is the assumption of spherical symmetry. In the case where the galaxy is flattened along the line-of-sight, its mass can be under-estimated if the system is assumed to be spherically symmetric \\citep{magorrian01}. As both X-ray and dynamical methods have their own intrinsic assumptions, the most robust constraints can be placed on the mass profiles of early-type galaxies when different methods are compared. Indeed, there is currently an emerging attempt to use different techniques in a complementary manner \\citep[][]{romanowsky08,churazov08,samurovic06,bridges06}; additionally, this approach improves our understanding of the systematics involved in each method. Recent work by \\citet{churazov08} explored in detail the comparison between X-ray and optically derived profiles for M87 and NGC 1399, finding agreement between the methods at the 10--20\\,\\% level when looking at the gravitational potential. However, both of these systems reside at the centres of clusters, M87 being the centre of Virgo, and NGC 1399 the centre of Fornax, and in both cases, the measurement of the potential is probing the cluster potential. In so-called `normal' elliptical galaxies, the situation is much less certain. For example, in the galaxy NGC 3379, \\citet{pellegrini06} require an outflow of the X-ray emitting gas to bring the X-ray results into agreement with the optically derived results. Only by the study of more systems with multiple approaches will we be able to reconcile the observed discrepancies. This is successful on a local scale, as individual GCs and/or planetary nebulae need to be resolved, limiting the distance to which these observations can be made. Investigating the wider properties of the dark matter halos of elliptical galaxies will require techniques such as stacked lensing \\citep[e.g.][]{sheldon04,hoekstra05,kleinheinrich06,koopmans06,mandelbaum06,ferreras08}. The layout of the paper is as follows. Section \\ref{sec:N4636} describes the basic properties of NGC~4636 and Section \\ref{sec:x-ray} describes our method for extracting high resolution mass profiles from \\Chandra\\ X-ray data. In Section \\ref{sec:results} we present our results and comparison to the GC analysis of \\citet{chakrabarty08}, and in Section \\ref{sec:discuss} we discuss the implications of our results. ", "conclusions": "We present an X-ray mass analysis of the early-type galaxy NGC~4636 using \\Chandra\\ data, under the assumptions of spherical symmetry and hydrostatic equilibrium. The integrity of the latter assumption has been questioned with reference to early-type galaxies \\citep{diehl06}, and it is because of the observed disturbances in the gas in NGC~4636 \\citep[e.g.][]{jones02,ohto03,osullivan05b} that we chose to study this object, in an effort to assess the impact of this assumption on the recovered mass profile. We find that the treatment of the abundance gradient in the X-ray analysis can significantly affect the recovered mass profile at all radii. We have compared the X-ray mass density profile with that recovered from a dynamical analysis of the system's globular clusters (GCs), presented by \\citet{chakrabarty08}. Inside 10\\,\\kpc, the dynamical mass estimate exceeds the X-ray mass estimate. The gas in this region is highly disturbed, and we postulate the cause of the disagreement to be a localised inflow of gas, or a contribution of non-thermal pressure support. The mass density profiles over the range $\\sim$10--30\\,\\kpc\\ are consistent within 1\\,$\\sigma$, indicating that even in this highly disturbed system, the recovered X-ray mass is consonant with that recovered from an independent method over intermediate radii. However, outside 30\\,\\kpc, the X-ray mass estimate exceeds the dynamical mass estimate, by a factor of 4--5 times at its greatest disagreement. Examining the anisotropy of the GCs, we find no statistical reason to reject our assumption of isotropy. The GC analysis is model-independent, so is not limited by the method, but the paucity of measured GC kinematics outside 7.5$^{\\prime}$ means that the success of this method at large radius is limited by the data. We test the assumption of hydrostatic equilibrium in our X-ray analysis, finding that local disturbances at large radii do not account for the observed discrepancy. At this radius, the group gas contribution is important in this system \\citep{osullivan05b}, and the overall state of the gas at this radius is uncertain. Mapping the X-ray properties to a larger radius using \\XMM\\ would help to model the group emission, but is beyond the scope of the current paper. The X-ray and dynamical mass analysis methods both indicate the need for a dark matter halo in this system, and provide a useful comparison within 30\\,\\kpc. It is through the comparison of independent approaches that the most robust constraints will be placed on the mass distribution of early-type galaxies, but we conclude that the limiting factors in such a comparison to large radius (outside 30\\,\\kpc) are data quality in the case of the GC kinematics, knowledge of the overall state of the gas as we reach the group regime in the case of the X-ray analysis, or a combination of the two." }, "0910/0910.0976_arXiv.txt": { "abstract": "Long-term trends in the solar spectral irradiance are important to determine the impact on Earth's climate. These long-term changes are thought to be caused mainly by changes in the surface area covered by small-scale magnetic elements. The direct measurement of the contrast to determine the impact of these small-scale magnetic elements is, however, limited to a few wavelengths, and is, even for space instruments, affected by scattered light and instrument defocus. In this work we calculate emergent intensities from 3-D simulations of solar magneto-convection and validate the outcome by comparing with observations from Hinode/SOT. In this manner we aim to construct the contrast at wavelengths ranging from the NUV to the FIR. ", "introduction": "Variations in the long-term spectral solar irradiance, i.e. the changes in the Sun's brightness at a certain wavelength, are significant for their effects on Earth's atmospheric temperature and composition. Magnetic fields concentrated in small structures influence these long-term changes, with increasing relevance towards the solar limb. Here, we compare observed and simulated intensity contrasts and investigate their behaviour from disk centre to the limb in 3D MHD simulations with different average vertical magnetic fields. ", "conclusions": "This analysis will allow us to determine the network and facular contrast as a function of magnetic flux, limb angle, and wavelength. We will then be able to remove the single free parameter in the SATIRE model (e.g. Krivova et al. 2003) and help improve irradiance reconstructions." }, "0910/0910.5886_arXiv.txt": { "abstract": "For axisymmetric models for coronal loops the relationship between the bifurcation points of magnetohydrodynamic (MHD) equilibrium sequences and the points of linear ideal MHD instability is investigated imposing line-tied boundary conditions. Using a well-studied example based on the Gold-Hoyle equilibrium, it is demonstrated that if the equilibrium sequence is calculated using the Grad-Shafranov equation, the instability corresponds to the second bifurcation point and not the first bifurcation point because the equilibrium boundary conditions allow for modes which are excluded from the linear ideal stability analysis. This is shown by calculating the bifurcating equilibrium branches and comparing the spatial structure of the solutions close to the bifurcation point with the spatial structure of the unstable mode. If the equilibrium sequence is calculated using Euler potentials the first bifurcation point of the Grad-Shafranov case is not found, and the first bifurcation point of the Euler potential description coincides with the ideal instability threshold. An explanation of this results in terms of linear bifurcation theory is given and the implications for the use of MHD equilibrium bifurcations to explain eruptive phenomena is briefly discussed. ", "introduction": "Magnetohydrodynamic instabilities of coronal loops are since a long time discussed as one of the main theoretical explanations for solar flares, in particular compact loop flares ({\\it e.g.} \\opencite{priest82}). Traditionally, investigations of MHD instabilities of coronal loops model these loops as straight cylindrical flux tubes of finite length with line tied boundary conditions at the `photospheric' ends of the flux tubes ({\\it e.g.} \\opencite{raadu72}; \\opencite{hood:priest79}, \\citeyear{hood:priest81}; \\opencite{einaudi:vanhoven83}; \\opencite{velli:etal90}). Such a set-up allows for a wide variety of relatively simple equilibrium configurations, hence explaining its popularity. The stability of equilibrium configurations of the above mentioned type has been studied for several decades using the methods of linear MHD stability analysis ({\\it e.g.} \\opencite{raadu72}; \\opencite{hood:priest79}, \\citeyear{hood:priest81}; \\opencite{einaudi:vanhoven83}; \\opencite{velli:etal90}; \\opencite{debruyne:hood89}, \\citeyear{debruyne:hood92}; \\opencite{mikic:etal90}; \\opencite{hood:etal94}; \\opencite{vanderlinden:hood98}, \\citeyear{vanderlinden:hood99}). In recent years the investigations have been extended into the nonlinear regime using large-scale MHD simulations ({\\it e.g.} \\opencite{longbottom:etal96}; \\opencite{baty:heyvaerts96}; \\opencite{baty97a}, \\citeyear{baty97b}, \\citeyear{baty00a}, \\citeyear{baty00b}; \\opencite{lionello:etal98}; \\opencite{arber:etal99}; \\opencite{gerrard:etal01}; \\opencite{browning:vanderlinden03}; \\opencite{browning:etal08}; \\opencite{hood:etal09}). In the present contribution we want to investigate the stability of line-tied coronal loop models from a different point of view. The flux tube equilibria used to model coronal loops all depend on one or more parameters representing quantities like the magnetic twist or the plasma beta. Many investigations study how the linear stability of the loops changes as one (or more) of these equilibrium parameters vary. The systematic variation of one or several parameters of an equilibrium defines an equilibrium sequence, and a point of linear instability should correspond to a bifurcation point of the equilibrium sequence and vice versa. It has to be kept in mind, however, that magnetostatic equilibria are usually calculated by solving a mathematically reduced set of equations. It is not at all clear whether there is really a one-to-one correspondence between points of linear instability and bifurcation points, in particular if line-tied boundary conditions are imposed as in models of coronal loops. In the present paper we shall investigate the question whether the points of linear instability of rotationally symmetric straight line-tied flux tubes have a one-to-one correspondence with the bifurcation points of equilibrium sequences. We shall use two different ways of calculating the equilibrium sequences, namely Grad-Shafranov theory and Euler potentials, and we shall, for simplicity, investigate only axisymmetric instabilities and bifurcations. A particularly well-studied equilibrium class \\cite{gold:hoyle60} will be used to carry out this investigation, mainly because results of linear stability investigations for this equilibrium class are readiliy available in the literature ({\\it e.g.} \\opencite{hood:priest79}, \\citeyear{hood:priest81}; \\opencite{mikic:etal90}; \\opencite{debruyne:hood92}). In Section \\ref{basic} the basic equilibrium theory and those parts of the theory of linear MHD stability needed in this paper are discussed. The following Section \\ref{numerics} presents a brief outline of the numerical method used to calculate the equilibrium sequences and to determine their bifurcation points and bifurcating branches. The results of these calculations are given in Section \\ref{results} and discussed in Section \\ref{discussion}. The paper closes with a summary in Section \\ref{summary}. ", "conclusions": "\\label{summary} We have investigated the relationship between MHD bifurcation and linear stability for a class of axisymmetric straight flux tubes under line-tying boundary conditions. For simplicity we only considered rotationally symmetric perturbations, allowing only for sausage modes. We have used two different ways of calculating the equilibrium sequences including bifurcating branches - one approach uses the Grad-Shafranov equation, the other approach uses Euler potentials. It turns out that only the Euler potential case shows a one-to-one correspondence between the first bifurcation point and the linear instability threshold for the sausage mode. The Grad-Shafranov case shows an additional bifurcation which does not correspond to the instability threshold under line-tying boundary conditions. This difference can be explained by the different constraints imposed on the bifurcating equilibrium branches in the Grad-Shafranov and the Euler potential cases. Furthermore, even though the second bifurcation point of the Grad-Shafranov case coincides with the first bifurcation point of the Euler potential case and the linear instability threshold, the structure of the bifurcation diagrams differ considerably between the Grad-Shafranov and the Euler potential case. The reason for this is not yet clear, but is probably also due to the difference in boundary conditions. In any case this difference has implications for the stability of the bifurcating equilibrium branches (see {\\it e.g.} \\opencite{iooss:joseph80}) and is therefore important to decide whether the system is able to find a new equilibrium (in the present case a new axisymmetric equilibrium) if one would consider an imaginary process driving the flux tube across the instability threshold. The present investigation is a preparation for studying equilibrium sequences of magnetic flux tubes and other solar magnetic structures together with their bifurcations in three dimensions. Preliminary steps have already been made (see {\\it e.g.} \\opencite{romeou:neukirch02}) and more detailed investigations are planned for the future. \\begin{acks} The authors thank Alan Hood for useful discussions. T. Neukirch acknowledges support by STFC and by the European Commission through the SOLAIRE Network (MTRN-CT-2006-035484). Z. Romeou gratefully acknowledges financial support provided through the European Community's Training and Mobility of Researchers Programme by a Marie-Curie Fellowship and through the European Community's Human Potential Programme under contract HPRN-CT-2000-00153, PLATON. The authors also acknowledge partial support by the British Council ARC Programme. \\end{acks} \\begin{appendix} Whereas the first bifurcating branch in the Grad-Shafranov case exists for values of $\\lambda$ which are both bigger and smaller than the $\\lambda$ at the bifurcation point, the second bifurcating branch exists only for $\\lambda$ smaller than the bifurcation $\\lambda$. This fact can be explained by using standard bifurcation theory to calculate the structure of the bifurcating branches close to the bifurcation points. The argument is actually independent of the form of the functions $p(A)$ and $b_\\phi(A)$. The qualitative structure of the bifurcation diagram will thus be the same even if $p(A)$ and $b_\\phi(A)$ are changed as long as the fundamental branch consists of solutions which depend only on the radial coordinate $r$. We start by writing the Grad-Shafranov equation in the form \\begin{equation} G(A,\\lambda,r)=-r\\nabla\\cdot\\left(\\frac{1}{r^2}\\nabla A\\right) - N(A,\\lambda,r) = 0, \\label{gsgeneral} \\end{equation} where the function $N(A,r,\\lambda)$ summarizes the nonlinear part of the Grad-Shafra\\-nov equation given by $p(A,\\lambda)$ and $b_\\phi(A,\\lambda)$. For the present paper $p$ and $b_\\phi$ are given by Equations (\\ref{ghpa}) and (\\ref{ghbpa}). For the following argument, however, the exact form of $N(A,r,\\lambda)$ is irrelevant, as long as it is analytic in $A$ and $\\lambda$ at the bifurcation points we want to investigate. We will not give here any details of the mathematical background which can be found for example in \\inlinecite{hesse:ks86} and\\inlinecite{hesse:kiessling87}. These papers treat slighly different bifurcation problems, but we will be using the same technique. Let $\\lambda^*$ be the value of $\\lambda$ at either of the bifurcation points and let $A_0=A_0(\\lambda^*)$ be the solution of Equation (\\ref{gsgeneral}) at the bifurcation point. To calculate the bifurcating branch we expand $\\lambda$ and $A$ as \\begin{eqnarray} \\lambda &=& \\sum_{k=0}^\\infty \\epsilon^k \\lambda_k ,\\label{lambdaexp} \\\\ A &=& \\sum_{k=0}^\\infty \\epsilon^k A_k , \\label{Aexp} \\end{eqnarray} where $\\lambda_0=\\lambda^*$ and $A_0$ as above. Since $G$ is analytic in both $A$ and $\\lambda$ for $\\lambda > 0$ we can expand Equation (\\ref{gsgeneral}) in a power series in $\\epsilon$: \\begin{equation} 0 = \\sum_{k=0}^\\infty \\frac{1}{k !}\\frac{d^k }{d\\epsilon^k} G(A(\\epsilon),\\lambda(\\epsilon),r )\\Big|_{\\epsilon=0} \\epsilon^k . \\end{equation} As each power of $\\epsilon$ must satisfy this equation independently we obtain \\begin{equation} \\frac{d^k }{d\\epsilon^k} G(A(\\epsilon),\\lambda(\\epsilon),r )\\Big|_{\\epsilon=0} = 0, \\qquad k=0,1,2,\\ldots\\; . \\label{Gexp} \\end{equation} Obviously, the lowest order equation \\begin{equation} G(A_0(\\lambda^*),\\lambda^*,r) = 0 \\end{equation} is just the Grad-Shafranov equation at the bifurcation point and therefore trivially satisfied. For the discussion of the higher order equations we first have to look at the boundary conditions the $A_k$ have to satisfy. The boundary condtion $A_b$ for $A(r,z,\\lambda)$ is given by the fundamental branch solution $A_b(r,z,\\lambda)=A_0(r,z,\\lambda)$ (the Gold-Hoyle solution in the present paper). Therefore we can extend $A_b$ into the domain. Using the expansion (\\ref{lambdaexp}) in $A_b(r,z,\\lambda(\\epsilon))$ we can see that the boundary condition each of the $A_k$ in Equation (\\ref{Aexp}) has to satisfy is given by \\begin{equation} A_b^{(k)} = \\frac{1}{k !}\\frac{d^k}{d \\epsilon^k} A_0(r,z,\\lambda(\\epsilon)) \\Big|_{\\epsilon=0}. \\end{equation} Note that $A_b^{(k)}$ satifies the same Equation (\\ref{Gexp}) as $A_k$. For $O(\\epsilon)$ we get from Equation (\\ref{Gexp}) \\begin{equation} G_A(A_0(\\lambda^*),\\lambda^*, r) A_1 + G_\\lambda(A_0(\\lambda^*),\\lambda^*, r) \\lambda_1 =0, \\end{equation} with \\begin{eqnarray} G_A A_1 &=& -r\\nabla\\cdot\\left(\\frac{1}{r^2}\\nabla A_1\\right) - \\frac{\\partial N}{\\partial A}(A_0,\\lambda^*,r) A_1 ,\\\\ G_\\lambda &=&\\frac{\\partial N}{\\partial \\lambda}(A_0,\\lambda^*,r). \\end{eqnarray} As mentioned above $A_b^{(1)}$ satisfies the same equation as $A_1$ and therefore the function \\begin{equation} A_1^\\prime = A_1 -A_b^{(1)} \\end{equation} satisfies $G_A A_1^\\prime =0$ or explicitely \\begin{equation} -r\\nabla\\cdot\\left(\\frac{1}{r^2}\\nabla A_1^\\prime\\right) - \\frac{\\partial N}{\\partial A}(A_0,\\lambda^*,r) A_1^\\prime =0 \\label{firstorder} \\end{equation} with $A_1^\\prime = 0$ on the boundaries. Since all coefficients of Equation (\\ref{firstorder}) depend only on $r$ its solution can be obtained by separation of variables with the general form of $A_1^\\prime$ being \\begin{equation} A_1^\\prime(r,z) = F_n(r) \\sin(n\\pi z/L), \\qquad n= 1,2,3,\\ldots \\;. \\label{a1modes} \\end{equation} The exact form of $F_n(r)$ is of no importance for the following argument. If we want to calculate $\\lambda_1$, we have to go to the next order ($O(\\epsilon^2)$) of the expansion, giving \\begin{eqnarray} & &-r\\nabla\\cdot\\left(\\frac{1}{r^2}\\nabla A_2\\right)-\\frac{\\partial N}{\\partial A} A_2 -\\frac{1}{2}\\frac{\\partial^2 N}{\\partial A^2} A_1^2 - \\frac{\\partial^2 N}{\\partial A \\partial \\lambda} A_1 \\lambda_1 \\nonumber \\\\ & &\\mbox{ \\hspace{5cm}}-\\frac{1}{2}\\frac{\\partial^2 N}{\\partial \\lambda^2} \\lambda_1^2 - \\frac{\\partial N}{\\partial \\lambda} \\lambda_2 =0 \\end{eqnarray} where all derivatives of $N(A,\\lambda,r)$ are evaluated at the bifurcation point ($\\epsilon=0$). Similarly to $A_b^{(1)}$ at $O(\\epsilon)$, $A_b^{(2)}$ satisfies the same equation as $A_2$. We define \\begin{equation} A_2^\\prime=A_2-A_b^{(2)} \\end{equation} which obeys the equation \\begin{equation} G_A A_2^\\prime = \\frac{1}{2}\\frac{\\partial^2 N}{\\partial A^2} ({A_1^\\prime}^2 + 2 A_b^{(1)} A_1^\\prime) + \\lambda_1\\frac{\\partial^2 N}{\\partial A \\partial \\lambda} A_1^\\prime . \\label{secondorder} \\end{equation} By Fredholm's alternative the right hand side of Equation (\\ref{secondorder}) has to be orthogonal to $A_1^\\prime$, {\\it i.e.} \\begin{equation} \\int_0^L\\int_0^{r_{max}}\\left( \\frac{1}{2}\\frac{\\partial^2 N}{\\partial A^2} ({A_1^\\prime}^2 + 2 A_b^{(1)} A_1^\\prime) + \\lambda_1\\frac{\\partial^2 N}{\\partial A \\partial \\lambda} A_1^\\prime \\right) A_1^\\prime rdrdz = 0 . \\label{fredha} \\end{equation} To proceed we assume in agreement with the Gold-Hoyle solution that the function $A_b^{(1)}$ has the form \\begin{equation} A_b^{(1)} = \\lambda_1 f_b(r) \\end{equation} where $f_b(r)$ is left unspecified here. Equation (\\ref{fredha}) can then be used to calculate $\\lambda_1$ in the form \\begin{eqnarray} \\lefteqn{ \\lambda_1 \\int_0^L\\int_0^{r_{max}}\\left( \\frac{\\partial^2 N}{\\partial A^2} f_b(r) + \\frac{\\partial^2 N}{\\partial A \\partial \\lambda}\\right) {A_1^\\prime}^2 r dr dz = } \\nonumber \\\\ & & \\mbox{\\hspace{4cm}}-\\int_0^L\\int_0^{r_{max}} \\frac{1}{2}\\frac{\\partial^2 N}{\\partial A^2} {A_1^\\prime}^3 r dr dz . \\label{l1equat} \\end{eqnarray} The double integral on the right hand side of Equation (\\ref{l1equat}) can be split into two separate integrations over $r$ and $z$, since the integrand depends on $z$ only through $A_1^\\prime$. As $A_1^\\prime$ has the form (\\ref{a1modes}), the integral over $z$ is given by \\begin{equation} \\int_0^L \\sin^3(n\\pi z/L) dz = -\\frac{L}{n\\pi}(\\cos n\\pi -1) +\\frac{L}{3n\\pi}(\\cos^3 n\\pi -1) , \\end{equation} which vanishes for all even $n$. As the integral on the left hand side of Equation (\\ref{l1equat}) is nonzero, this implies that for even $n$ (and in particular for $n=2$) $\\lambda_1$ vanishes. The bifurcation at bifurcation points with modes having even $n$ is therefore quadratic. This explains the structure of the bifurcation diagram in Figure \\ref{bifdiaggs7}, because the first bifurcation obviously corresponds to the $A_1^\\prime$ for $n=1$, whereas the second bifurcation corresponds to $n=2$. Therefore the structure of the first branch close to the bifurcation point is given by \\begin{eqnarray} \\lambda & = &\\lambda^* + \\epsilon \\lambda_1 + \\ldots, \\label{lambdabif1} \\\\ A & = & A_0 (r,z,\\lambda^*) + \\epsilon A_1(r,z,\\lambda^*) + \\ldots.\\label{Abif1} \\end{eqnarray} The slope of the bifurcating branch at the bifurcation point is determined by $\\lambda_1 \\ne 0$ in this case and it is obvious that the bifurcating branch exists for both $\\lambda > \\lambda^*$ ($\\epsilon \\lambda_1 > 0$) and $\\lambda < \\lambda^*$ ($\\epsilon \\lambda_1 < 0$). Close to the second bifurcation point we have \\begin{eqnarray} \\lambda & = &\\lambda^* + \\frac{1}{2}\\epsilon^2 \\lambda_2 + \\ldots, \\label{lambdabif2}\\\\ A & = & A_0 (r,z,\\lambda^*) + \\epsilon A_1(r,z,\\lambda^*) + \\ldots, \\label{Abif2} \\end{eqnarray} because here $\\lambda_1$ vanishes. As the correction to $\\lambda^*$ depends quadratically on $\\epsilon$, positive and negative $\\epsilon$ give the same value of $\\lambda$. This implies that the second bifurcating branch actually consists of {\\em two} branches, one for positive and one for negative $\\epsilon$. Since a change of sign of $\\epsilon$ in Equation (\\ref{Abif2}) corresponds to a simple mirroring of the $\\sin(2\\pi z/L)$ function at the point $z=L/2$, the two branches have exactly the same energies. We remark that since $\\lambda_1=0$ in this case $A_1 = A_1^\\prime$ as the boundary contribution to $A_1^\\prime$ vanishes. The numerical calculations corroborate these results as the same second bifurcation branch is found by the code starting both with negative and positive $\\epsilon$. The only difference between the calculations is the mirroring of the $z$-dependence of $A$ along the bifurcating branch. \\end{appendix}" }, "0910/0910.5248_arXiv.txt": { "abstract": "Microgauss magnetic fields are observed in all galaxies at low and high redshifts. The origin of these intense magnetic fields is a challenging question in astrophysics. We show here that the natural plasma fluctuations in the primordial universe (assumed to be random), predicted by the Fluctuation-Dissipation-Theorem, predicts $\\sim 0.034~ \\mu G$ fields over $\\sim 0.3$ kpc regions in galaxies. If the dipole magnetic fields predicted by the Fluctuation-Dissipation-Theorem are not completely random, microgauss fields over regions $\\gtrsim 0.34$ kpc are easily obtained. The model is thus a strong candidate for resolving the problem of the origin of magnetic fields in $\\lesssim 10^{9}$ years in high redshift galaxies. ", "introduction": "The origin of large-scale cosmic magnetic fields in galaxies and protogalaxies remains a challenging problem in astrophysics \\citep{zwe97, kul08, raf08, wid02}. There have been many attempts to explain the origin of cosmic magnetic fields. One of the first popular astrophysical theories to create seed fields was the Biermann mechanism \\citep{bie50}. It has been suggested that this mechanism acts in diverse astrophysical systems, such as large scale structure formation \\citep{pee67,ree72,was78}, cosmological ionizing fronts \\citep{gne00}, star formation and supernova explosions \\citep{mir98, han05}. \\citet{ryu08} made simulations showing that cosmological shocks can create average magnetic fields of a few $\\mu G$ inside cluster/groups, $ \\sim 0.1 ~\\mu G$ around clusters/groups, and $\\sim 10~nG$ in filaments. \\citet{med06} showed that magnetic fields can be produced by collisionless shocks in galaxy clusters and in the intercluster medium (ICM) during large scale structure formation. \\citet{ars09} studied the evolution of magnetic fields in galaxies coupled with hierarchical structure formation. \\citet{ich06} investigated second-order couplings between photons and electrons as a possible origin of magnetic fields on cosmological scales before the epoch of recombination. The creation of early magnetic fields generated by cosmological perturbations have also been investigated \\citep{tak05, tak06, cla01, mae09}. In our galaxy, the magnetic field is coherent over kpc scales with alternating directions in the arm and inter-arm regions (e.g.,\\citet{kro94, han08}). Such alternations are expected for magnetic fields of primordial origin \\citep{gra01}. Various observations put upper limits on the intensity of a homogeneous primordial magnetic field. Observations of the small-scale cosmic microwave background (CMB) anisotropy yield an upper comoving limit of $4.7 ~nG$ for a homogeneous primordial field \\citep{yam06}. Reionization of the Universe puts upper limits of $0.7-3~ nG$ for a homogeneous primordial field, depending on the assumptions of the stellar population that is responsible for reionizing the Universe \\citep{sch08}. Another upper limit for a homogenous primordial magnetic field is the magnetic Jeans mass $\\sim 10^{10} M_{\\odot} ~(B/3nG)^{3}$ \\citep{sub98, set05}. Thus, if we are investigating the collapse of a $\\sim 10^{7} M_{\\odot}$ protogalaxy, the homogeneous primordial magnetic field must be $< 0.3~ nG$ in order for collapse to occur. Galactic magnetic fields have been suggested to have evolved in three main stages. In the first stage, seed fields were embedded in the protogalaxy. They may have had a primordial origin, as suggested in this paper. Another possibility is that the seed fields could have been injected into the protogalaxies by AGN jets, radio lobes, supernovas, or a combination of the above. Still another possibility is that the seed fields may have been created by the Biermann battery during the formation of the protogalaxy. In the second stage, the seed fields were amplified by compression, shearing flows, turbulent flows, magneto-rotational instabilities, dynamos or by a combination of the above. In the last stage magnetic fields were ordered by a large scale dynamo \\citep{beck06}. \\citet{ryu08} investigated the amplification of magnetic fields due to turbulent vorticity created at cosmological shocks during the formation of large scale structures. A given vorticity $\\omega$ can be characterized by a characteristic velocity $V_{c}$ over a characteristic distance $L_{c}$. \\citeauthor{ryu08} found that $\\omega$ typically is \\begin{equation} \\omega \\sim 1-3 \\times 10^{-16} s^{-1}, \\end{equation} which corresponds to 10-30 turnovers in the age of the universe. They investigated $L_{c} > $ 1 Mpc $h^{-1}$. We investigate $L_{c} \\simeq$ 200 kpc $h^{-1}$ in protogalaxies for a similar vorticity. We show that a seed field $0.003 ~nG$ over a comoving $2~ kpc$ region at \\emph{z} $\\sim 10$, predicted by the Fluctuation-Dissipation Theorem \\citep{raf08}, amplified by the small scale dynamo is a good candidate for the origin of magnetic fields in galaxies. \\citet{sub97, sub99} and \\citet{bra00} derived the non-linear evolution equations for the magnetic correlations. We use their formulation for the small scale dynamo and solve the nonlinear equations numerically. In \\S II, we review the creation of magnetic fields due to electromagnetic fluctuations in hot dense equilibrium primordial plasmas, as described in our previous work \\citep{raf08}. In \\S III, we discuss the small scale dynamo and in \\S IV, the important parameters of the plasma to be used in the calculations. In \\S V, we present our results and in \\S VI our conclusions. ", "conclusions": "It was shown previously that the magnetic fields, created immediately after the quark-hadron transition, produce relatively intense magnetic dipole fields on small scales at $\\emph{z} \\sim 10$ \\citep{raf08}. We show here that the predicted seed fields of size $\\sim 2$ kpc and intensity $0.003~ nG$ at \\emph{z} $\\sim 10$ can be amplified by a small scale dynamo in protogalaxies to intensities close to observed values. In the small scale dynamo studied, we use the turbulent spectrum given by \\citet{sub99}. The characteristic velocity $V_{c}$ and length $L_{c}$, used in the expression for the vorticity $V_{c}/L_{c}$, are $V_{c} \\simeq 10^{7} cm/s$ and $L_{c} \\simeq$ 200 kpc. This vorticity is comparable to that found by \\citet{ryu08}, studying the formation of large scale structures. The length $L_{c} \\simeq 200$ kpc used is a characteristic size of a protogalactic cloud. The turbulent spectrum used simulates Kolmogorov turbulence \\citep{vai82}. From our Figs. 1 and 2, we find that $M_{L}(\\sim B^{2})$ increases from $\\sim 10^{-23} ~G^{2}$ (corresponding to a magnetic field $B\\sim 3\\times 10^{-12} ~G$ over a region $L\\sim $ 2 kpc) to $M_{L} \\sim 10^{18}~G^{2}$ (corresponding to a field $\\sim 10^{-9}~G$ over a region $L\\sim 2$ kpc) in $10^{9}$ years. This corresponds to a $\\sim 6$ e-fold amplification of $B$ in a relatively short time. Collapsing to form galaxies at redshift $\\emph{z}\\sim10$, the density increases by a factor of $\\sim 200$ and the magnetic fields are amplified by a factor of $\\sim 34$. This predicts $0.03~ \\mu G$ fields over 0.34 kpc regions in galaxies. If the dipole magnetic fields predicted by the Fluctuation-Dissipation Theorem are not completely random, microgauss fields over regions $> 0.34$ kpc are easily obtained. The model studied is thus a strong candidate to explain the $\\mu G$ fields observed in high redshift galaxies. \\begin{figure} \\centering \\includegraphics[scale=0.55]{f1.eps} \\caption{Values of $M_{L}(G^{2}$) as a function of t (years) and r. Solid black line has the reference values: $M_{L}(r,0)= 10^{-11} (0.1pc/r)^{3}~G^{2}$, $L_{c} = 200 ~kpc$, r = 3 kpc, and $V_{c}=10^{7}~ cm/s$ in Eqs. (36)-(38). Dashed red line is for $r= 4 ~kpc$, and the dotted blue line for $r = ~5 kpc$.} \\label{f1} \\end{figure} \\begin{figure} \\centering \\includegraphics[scale=0.55]{f2.eps} \\caption{Values of $M_{L}(G^{2}$) as a function of t (years), varying $V_{c}$. Solid black line has the reference values in Fig. \\ref{f1}. Dashed red line is for $V_{c}= 8\\times10^{6}~cm/s$. Dotted blue line is for $V_{c}= 6\\times10^{6}~ cm/s$.} \\label{f3} \\end{figure} \\begin{figure} \\centering \\includegraphics[scale=0.55]{f3.eps} \\caption{Values of B(G) as a function of t (years) and r(kpc) for reference values of Fig. 1.} \\label{f33} \\end{figure}" }, "0910/0910.4954_arXiv.txt": { "abstract": "{The solar rotation profile is conical rather than cylindrical as one could expect from classical rotating fluid dynamics (e.g. Taylor-Proudman theorem). Thermal coupling to the tachocline, baroclinic effects and latitudinal transport of heat have been advocated to explain this peculiar state of rotation.} {To test the validity of thermal wind balance in the solar convection zone using helioseismic inversions for both the angular velocity and fluctuations in entropy and temperature.} {Entropy and temperature fluctuations obtained from 3-D hydrodynamical numerical simulations of the solar convection zone are compared with solar profiles obtained from helioseismic inversions.} {The temperature and entropy fluctuations in 3-D numerical simulations have smaller amplitude in the bulk of the solar convection zone than those found from seismic inversions. Seismic inversion find variations of temperature from about 1 K at the surface up to 100 K at the base of the convection zone while in 3-D simulations they are of order 10 K throughout the convection zone up to 0.96 $R_{\\odot}$. In 3-D simulations, baroclinic effects are found to be important to tilt the isocontours of $\\Omega$ away from a cylindrical profile in most of the convection zone helped by Reynolds and viscous stresses at some locations. By contrast the baroclinic effect inverted by helioseismology are much larger than what is required to yield the observed angular velocity profile.} {The solar convection does not appear to be in strict thermal wind balance, Reynolds stresses must play a dominant role in setting not only the equatorial acceleration but also the observed conical angular velocity profile.} ", "introduction": "Helioseismic data from the Global Oscillation Network Group (GONG) and the Michelson Doppler Imager (MDI) have been used to infer the rotation profile in the solar interior (e.g., Thompson et al.~1996; Schou et al.~1998). The inversion results show that isocontours of the differential rotation $\\Omega(r,\\theta)$ are conical at mid-latitude, rather than cylindrical as was expected from early numerical simulations (e.g., Glatzmaier \\& Gilman 1982; Gilman \\& Miller 1986). More recent theoretical work (Durney 1999; Kitchatinov \\& Rudiger 1995; Brun \\& Toomre 2002 (hereafter BT02); Rempel 2005; Miesch et al.~2006 (hereafter MBT06); Brun \\& Rempel 2008; Balbus et al. 2009) indicate that in order to break the Taylor-Proudman constraint of cylindrical $\\Omega$, the Sun must either have a systematic latitudinal heat transfer in its convection zone or thermal forcing from the tachocline or most likely both. This is due to the so-called thermal wind balance (Pedlosky 1987), i.e., the existence in the solar convection zone of latitudinal entropy (or temperature) variation due to baroclinic effect can result in a rotation state that breaks the Taylor-Proudman constraint. Such latitudinal variations of the thermal properties at the solar surface have been looked for observationally by several groups since the late 60's (e.g., Dicke \\& Goldenberg 1967; Altroch \\& Canfield 1972; Koutchmy et al.~1977; Kuhn et al.~1985, 1998, Rast et al.~2008; to cite only a few). This is a difficult task since one has to correct for limb darkening effect, photospheric magnetic activity, instrument bias and many other subtle effects to extract a relatively weak signal (see Rast et al.~2008). In most cases a temperature contrast of a few degree K is found from equator to pole at the surface, the pole being warmer. In some observations a minimum at mid latitude with a warm equator and hotter polar regions is also found. The warm polar regions and cool equatorial region pattern is also found in 3-D simulation of the solar convection zone with temperature variation slightly larger (i.e., of order 10 K; BT02, MBT06). At the surface a banded structure of the temperature field (warm-cool-hot) is also found in 3-D simulation of global scale convection. While very useful and instructive, most observations are confined to the solar surface and lack the information on the deep thermal structure of the solar convection zone which is key to characterise the dynamics of the deep solar convection zone. One way to remedy that limitation is to rely on helioseismic inversions that allows us to probe deeper in the Sun and to use 3-D global simulations of the solar convection zone to guide our physical understanding. Indeed, helioseismic inversions can give us the rotation rate, as well as the sound speed and density in the solar interior as a function of radius and latitude. Inside the convection zone the chemical composition is uniform and if we know the equation of state it is possible to determine other thermodynamic quantities like the temperature and entropy from the sound speed and density. Although, there may be some uncertainty in the equation of state, the OPAL equation of state (Rogers et al.~1996; Rogers \\& Nayafonov 2002) is quite close to the equation of state of solar material (e.g., Basu \\& Antia 1995; Basu \\& \\jcd~1997). Thus, in this work we use the OPAL equation of state to calculate the perturbations in entropy and temperature and assess how well a strict thermal wind balance is established in the solar convective envelope. To achieve this goal we make use of 2-D inversions of $\\Omega, S, T$, using the GONG and MDI data for the full solar cycle 23 and analyse our findings using 3-D simulations obtained with the ASH (anelastic spherical harmonic) code (BT02; MBT06; Miesch et al.~2008) supported by theoretical considerations on the thermal wind balance and vorticity equations. The rest of the paper is organised as follows: in Sect.~2 we describe the data and technique used in this work while the results for the temperature and entropy inversions are described in \\S~3 along with those of 3-D simulations. In \\S~4 we discuss at length the thermal wind balance and its generalisation and interpret our seismic inversion with 3-D simulation of global scale convection. Finally, in \\S~5 we put our results in perspective and conclude. \\begin{figure*} \\centerline{\\resizebox{0.65\\figwidth}{!}{\\includegraphics{f1a.eps} }\\hfill \\resizebox{0.65\\figwidth}{!}{\\includegraphics{f1b.eps} }\\hfill \\resizebox{0.68\\figwidth}{!}{\\includegraphics{f1c.eps} }} \\caption{The aspherical component of temperature fluctuation, $\\delta T$ obtained from the temporally averaged GONG (left panel) and MDI (middle panel) data. The right panel shows the cuts at constant latitude of $\\delta T$ obtained from MDI data along with $1\\sigma$ error estimates shown by dotted lines. All curves appear to merge at $r=R_\\odot$, because $\\delta T$ is of order of 1 K in that region.} \\label{T_obs} \\end{figure*} \\begin{figure*} \\centerline{\\resizebox{0.65\\figwidth}{!}{\\includegraphics{f2a.eps} }\\hfill \\resizebox{0.65\\figwidth}{!}{\\includegraphics{f2b.eps} }\\hfill \\resizebox{0.65\\figwidth}{!}{\\includegraphics{f2c.eps} }} \\caption{The aspherical component of entropy fluctuation, $\\delta S$ obtained from the temporally averaged GONG (left panel) and MDI (middle panel) data. The right panel shows the cuts at constant latitude of $\\delta S$ obtained from MDI data along with $1\\sigma$ error estimates shown by dotted lines.} \\label{S_obs} \\end{figure*} ", "conclusions": "What can be the source of the disagreement between the inverted baroclinic contribution and the $z$ derivative of the angular velocity (i.e equations 8, or 9 and 10)? The first and easiest solution is that the inversion of the thermal quantities lack the necessary accuracy and given the increase by two orders of magnitude of the background temperature and density with depth, we end up with variations that are too large. The source of discrepancy will then be due to an overestimation of $\\delta T$ and $\\delta S$. It is not easy to decide if these inverted thermal fluctuations are too large or if the simulations (both 2-D and 3-D) underestimates the fluctuations realised in the Sun, because for instance of their limited Reynolds number. We must thus also consider the possibility that these large thermal perturbations are genuine. If this is indeed the case we need to see how we could resolve the discrepancy between the seismically inverted LHS and RHS of equation 8. As stated in section \\ref{theory}, to obtain a strict thermal wind balance as expressed in equation 8, one need to make a certain number of assumptions: adiabaticity, weak Rossby number, negligible compressibility, viscous and Reynolds stresses, stationarity. Further by considering only the hydrodynamic contributions we have omitted those associated with Maxwell stresses that are certainly present in the magnetic Sun. Let's however assume that the Maxwell stresses are not the source of the large observed discrepancy. We are confident that this is the case because we have formed temporal averages over a maximum and minimum period of activity and the differences between the two periods are about 10 times smaller that what it would required if all the sources of discrepancy were coming from the Maxwell stresses alone. We nevertheless intend to make a more systematic study of the departure of the strict thermal wind balance linked to magnetic effects (i.e. via the so called magnetic wind) by analysing the solar cycle 23 in details and by comparing with dynamo simulations of the solar convection (Brun et al.~2004). We must thus question the validity of the other hypothesis made in deriving equation 8. Clearly it is justified given the very low microscopic value of the solar kinematic viscosity to consider that the viscous terms do not contribute much. This is clearly not the case in the 3-D models where near the surface they are major contributors to the overall balance (see Figure \\ref{TW_theo}, middle panel of the bottom row), but this is due to our large effective viscosity. Assuming adiabaticity is certainly reasonable in most of the convection zone but clearly not near the surface. Since we are mostly interested in understanding the bulk dynamics of the solar convection zone, this term is indeed very small. The choice of low Rossby number that allows us to neglect $\\omega$ over $2\\Omega_0$ the planetary vorticity is certainly not justified at all scales of the turbulent velocity spectra, in particular for those scales much smaller than the Rossby radius of deformation (Pedlosky 1987). In the Sun the large range of convection scales certainly undergo different dynamics depending on how sensitive they are to the Coriolis force. The subtle angular momentum and heat redistribution realised in the Sun is in part captured in our 3-D models. We can thus analyse if the Reynolds stresses associated with the turbulent motion indeed play a central role. As discussed in detail in Brun \\& Toomre (2002) and in \\S 4 we know that it is indeed the case in our numerical simulations (see Figure \\ref{TW_theo}, middle and right panel of the top row) even though our simulation do not possess a Reynolds number and a degree of turbulence as high as that in the Sun. We can thus expect, given the very large Reynolds number of the solar convection zone, that Reynolds stresses must play a central role in the Sun in shaping the differential rotation profile and that they somehow in part compensate the baroclinic contribution to yield the observed profile of angular velocity. This is a very interesting results, since it indicates that the differential rotation is of dynamical origin for its equatorial acceleration (as revealed by studying angular momentum transport in our simulations as in BT02 or Miesch et al.~2008) but also most certainly for its shape with Reynolds stresses helping or opposing in some regions the baroclinic effects to break Taylor-Proudman constraint. Of course this conclusion only holds if the inverted large thermal fluctuations are real." }, "0910/0910.5025_arXiv.txt": { "abstract": "The most X-ray luminous cluster known, RXJ1347-1145 ($z=0.45$), has been the object of extensive study across the electromagnetic spectrum. We have imaged the Sunyaev-Zel'dovich Effect (SZE) at 90 GHz ($\\lambda = 3.3$ mm) in RXJ1347-1145 at $10''$ resolution with the 64-pixel MUSTANG bolometer array on the Green Bank Telescope (GBT), confirming a previously reported strong, localized enhancement of the SZE $20''$ to the South-East of the center of X-ray emission. This enhancement of the SZE has been interpreted as shock-heated ($> 20 \\, {\\rm keV}$) gas caused by an ongoing major (low mass-ratio) merger event. Our data support this interpretation. We also detect a pronounced asymmetry in the projected cluster pressure profile, with the pressure just east of the cluster core $\\sim 1.6 \\times$ higher than just to the west. This is the highest resolution image of the SZE made to date. ", "introduction": "The rich cluster RXJ1347-1145 ($z=0.45$) is the most X-ray luminous galaxy cluster known \\citep{schindler95,schindler97,allen02} and has been the object of extensive study at radio, millimeter, submillimeter, optical and X-ray wavelengths \\citep{kitayama04,komatsu01,Gitti07,allen02,schindler97,pointecouteau99,ota08, cohen02,bradac08,miranda08}. Discovered in the ROSAT All-Sky Survey, RXJ1347-1145 was originally thought to be a dynamically old, relaxed system \\citep{schindler95,schindler97} based on its smooth, strongly-peaked X-ray morphology--- a prototypical relaxed ``cooling-flow'' cluster. The NOBA 7 bolometer system on the 45-meter Nobeyama telescope \\citep{kitayama04,komatsu01} has made high-resolution observations ($13''$ FWHM, smoothed to $\\sim 19''$ in the presented map) of the Sunyaev-Zel'dovich effect (SZE) at 150 GHz which indicate a strong enhancement of the SZ effect $20'' \\, (170 \\, {\\rm kpc})$ to the south-east of the peak of the X-ray emission, however. Hints of this asymmetry had been seen in earlier, lower resolution measurements with the Diabolo $2.1$~mm photometer on the IRAM 30-m \\citep{pointecouteau99}. The enhancement has been interpreted as being due to hot ($T_e > 20 \\, {\\rm keV}$) gas which is more difficult to detect using X-rays than cooler gas is, owing to the lower responsivities of imaging X-ray telescopes such as Chandra and XMM at energies above $\\sim 10 \\, {\\rm keV}$. In contrast, the SZE intensity is proportional to $T_e$ up to arbitrarily high temperatures, aside from relativistic corrections which are weak at 90 GHz, so such hot gas stands out. The feature is consistent with the presence of a large substructure of gas in the intra-cluster medium (ICM) shock-heated by a merger, as is seen in the ``Bullet Cluster'' 1E0657-56 \\citep{markevitch02}; this interpretaion has been supported by more recent observations \\citep[e.g.][]{allen02,ota08}. {\\it Thus, rather than being an example of a hydrostatic, relaxed system, high-resolution SZE observations suggest that the observed properties of the ICM in RXJ1347-1145 are strongly affected by an ongoing merger.} This is a striking cautionary tale for ongoing blind SZE surveys \\citep{carlstrom02}, for which useful X-ray data will be difficult or impossible to obtain for many high-$z$ systems, as well as a sign that our current understanding of nearby, well-studied X-ray clusters may be dramatically incomplete. Reports \\citep{komatsu01,pointecouteau01,kitayama04} of a strong enhancement of the SZE away from the cluster center are based on relatively low-resolution images compared to the size of the offsets and features involved. SZE images at lower frequencies also show show substantial offsets between the peak of X-ray and SZE emission; for instance, the 21 GHz \\citep{komatsu01} and SZE peak is $\\sim 20''$ to the SE of the X-ray peak, and the 30 GHz \\citep{reese02} SZE peak is $\\sim 13''$ to the SE of the X-ray peak. The situation is further complicated by the presence of a radio source in the center of the cluster. We have sought to test these claims, and to begin to untangle the astrophysics of this interesting system, with higher resolution imaging at a complementary frequency. In this paper we present the highest angular resolution image of the SZE yet made. We observed RXJ1347-1145 with the MUSTANG 90 GHz bolometer array on the Robert C. Byrd Green Bank Telescope (GBT). At the redshift of the cluster (and assuming $\\Omega_{\\Lambda}= 0.3,\\Omega_{tot}=1, h=0.73$) the GBT+MUSTANG $9''$ beam corresponds to a projected length of 54 kpc. The observations are described in \\S~\\ref{sec:obs} and the data reduction in \\S~\\ref{sec:reduc}. Our interpretation and conclusions are presented in \\S~\\ref{sec:concl}. ", "conclusions": "\\label{sec:concl} \\subsection{Comparison with Previous SZE Observations} Figure~\\ref{fig:Noba} presents a direct comparison of the MUSTANG and NOBA results in units of main-beam averaged Compton y parameter. For a more accurate comparison, we downgrade the resolution and pixelscale of the MUSTANG map to match that of NOBA (13'' FWHM on a 5'' pixel grid). The overall agreement between the maps is excellent, in particular as regards the amplitude and morphology of the local enhancement of the SZE south-east of the cluster core. The largest discrepancy is south west of the cluster, where NOBA shows a $3\\sigma$ compact decrement which is absent from the MUSTANG data. Considering the low and uniform X-ray surface brightness in the vicinity of this discrepancy (see Figure~\\ref{fig:composite}) and the higher angular resolution and lower noise of the MUSTANG data, it is likely that this feature is a spurious artifact in the NOBA map. Both datasets also show a ridge extending north from the shock front on the eastern side of the cluster. In the 150 GHz map the feature is of marginal significance ($1-2\\sigma$); interestingly, it is clearly visible in the 350 GHz SZE increment map but K04 dismiss it due to the possibility of confusing dust emission from the nearby galaxies. \\subsection{Empirical Model of the SZE in RXJ1347-1145} We construct a simple empirical model for the cluster SZE assuming the isothermal $\\beta$-model of \\citet{schindler97} normalized by the SZE measurement of \\citet{reese02} and \\citet{kitayama04} to describe the bulk cluster emission. We add a 5 mJy point source in the cluster core, coincident with the peak of the $\\beta$-model, and two Gaussian components in integrated pressure, one south-east and one almost directly east of the cluster center. In comparing to our 90 GHz data, we use the relatavistic correction of \\citet{sazonov98}, assuming $kT = 25 \\, {\\rm keV}$ (which reduces the amplitude of the decrement by 15\\%) for the Gaussian components and $kT = 10 \\, {\\rm keV}$ for the bulk component. The parameters chosen (two Gaussian widths for each component, a position, a peak surface brightness, and a position angle) are shown in Table~\\ref{tbl:szmodel}. The resulting sky image is convolved with our PSF (\\S~\\ref{sec:beam}) and transfer function (\\S~\\ref{sec:sims}). We find that this provides a good match to the data (Figure~\\ref{fig:szmodel}). The peak comptonization at $10''$ Gaussian resolution is $3.9 \\times 10^{-4}$ on the eastern ridge and $6.0 \\times 10^{-4}$ on the region identified as a shock by Komatsu et al. When convolved to $19''$ FWHM (NOBA) resolution, we find $\\Delta y = 3.9 \\times 10^{-4}$, close to their observed value $\\Delta y = 4.1 \\times 10^{-4}$. The intent of this static, phenomenological model is simply to provide a description of the observed high angular-resolution SZE and a direct comparison of NOBA and MUSTANG results. Work is underway which will allow quantitatively determining the best fit physical model by simulataneously fitting datasets at multiple wavelengths using a Monte-Carlo Markov Chain. This work is beyond the scope of this paper and will be presented in a follow up publication. \\begin{table*} \\begin{center} \\begin{tabular}{llll} Component & Amplitude & Offset & Notes \\\\ & [$y/10^{-3}$] & [$''$] & \\\\ \\hline $\\beta$-model & $1.0$ & $0,0$ & $\\theta_c = 10''$, $\\beta=0.60$ \\\\ Shock & $1.6$ & -14, 14 & $\\sigma_1=8''$, $\\sigma_2=2''$, ${\\rm P.A.}=45^{\\circ}$ \\\\ Ridge & $1.0$ & 10, 14 & $\\sigma_1=8''$, $\\sigma_2=2''$, ${\\rm P.A.}=-15^{\\circ}$ \\\\ \\hline \\end{tabular} \\caption{Note: Offset is (north,east) of peak X-ray position} \\label{tbl:szmodel} \\end{center} \\end{table*} \\subsection{Multi-wavelength Phenomenology} Our data show an SZE decrement with an overall significance of $5.4 \\sigma$. At the center of the cluster, coincident with the peak of X-ray emission and the brightest cluster galaxy (BCG), there is an unresolved $5 \\mJy$ radio source. This flux density is consistent with the 90 GHz flux density presented in \\citet{pointecouteau01}, as well as what is expected from a power law extrapolation of $1.4$ GHz and 30 GHz measurements \\citep{NVSS,reese02}. A strong, localized SZE decrement can be seen $20''$ to the south-east of the center of X-ray emission and clearly separated from the cluster center. Our data also indicate a high-pressure ridge immediately to the east of the cluster center. K04 tentatively attribute the south-east enhancement to a substructure of gas $240 \\pm 183 \\, {\\rm kpc}$ in length along the line of sight, at a density (assumed uniform) of $(1.4 \\pm 0.59) \\times 10^{-2} \\, {\\rm cm^{-3}}$ and with a temperature $T_e = 28.5 \\pm 7.3 \\, {\\rm keV}$. Recent X-ray spectral measurements \\citep{ota08} with SUZAKU also indicate the presence of hot gas in the south-east region ($T_e = 25.1^{+6.1}_{-4.5} \\ ^{+6.9}_{-9.5} \\, {\\rm keV}$ with statistical and systematic errors, respectively, at 90\\% confidence level). \\citet{allen02} have reported that the slight enhancement of softer X-ray emission in this region seen by Chandra is consistent with the presence of a small substructure of hot, shocked gas. \\citet{kitayama04} attribute the hot gas to an ongoing merger in the plane of the sky. The merger hypothesis is supported by optical data, in particular, the presence of a second massive elliptical $\\sim 20''$ directly to the east of the BCG that coincides with the center of X-ray emission (and with the radio point source). Furthermore the density and temperature of the hot substructure indicate that it is substantially overpressured compared to the surrounding ICM. Assuming a sound speed of $1600 \\, {\\rm km/sec}$ this overpressure region should relax into the surrounding ICM on a timescale $\\sim 0.1 \\, {\\rm Gyr}$, again arguing for an ongoing merger. Our data support this merger scenario. To put them in context, Figure~\\ref{fig:composite} shows a composite image with archival Chandra and HST data, and the weak + strong lensing mass map of \\citet{bradac08}. We propose that the data are best explained by a merger occuring in or near the plane of the plane of the sky. The left-hand (``B'') cluster, having fallen in from the south-{\\it west}, has just passed closest approach and is hooking around to the north-west. As the clusters merge shock forms, heating the gas in the wake of its passage. As argued by \\cite{kitayama04}, and seen in simulations \\citep{takizawa99}, the clusters must have masses within a factor of 2 or 3 of equality and a substantial ($\\sim 4000 \\, {\\rm km/sec}$) relative velocity in order to produce the high observed plasma temperatures, $T_e > 20 \\, {\\rm keV}$. This merger geometry is consistent with the lack of structure in the line-of-sight cluster member galaxies' velocities \\citep{cohen02}. The primary (right-hand, ``A'') cluster contains significant cold and cooling gas in its core (a ``cooling flow''). Such gas is seen to be quite robust in simulated major cluster mergers \\citep{gomez02,poole08}. Even in cases where the cooling flow is finally disrupted by the encounter, \\cite{gomez02} find a delay of $1-2$ GYr between the initial core encounter and the collapse of the cooling flow. The existence of a strong cooling flow, therefore, does not argue against a major merger in this case. More detailed simulations could shed further light on this interesting system. \\subsection{Broad Implications} Since calibrating SZ observable - mass relationships is vital to understand the implications of ongoing SZE surveys, it is important to understand the mechanism by which a substantial portion of the ICM can be heated so dramatically, and how this energy is distributed through the ICM over time. Observations of cold fronts in other clusters \\citep[e.g.][]{v01} have shown that energy transport processes in the ICM are substantially inhibited, perhaps by magnetic fields. It is distinctly possible, then, that once heated by shocks, very hot phases would persist. We have estimated the magnitude of the bias in an arcminute-resolution Compton $y$ measurement that is introduced by a hot gas phase by convolving the two gaussian components of the SZE model in Table~\\ref{tbl:szmodel} with a $1'$ FWHM Gaussian beam, typical of SZ survey telescopes such as ACT \\citep{actref,actclus} or SPT \\citep{sptref,sptcat}. Compared with the bulk emission component, also convolved with a $1'$ beam, the small-scale features are a 10\\% effect. While relatively modest this is a systematic bias in the Compton $y$ parameter which, if not properly accounted for, would result in a $20\\%$ overestimate in distances (underestimate in $H_0$) derived from a comparison of the SZE and X-ray data which did not allow for the presence of the hot gas component. To assess the impact on the {\\it scatter} in $M-y$, a larger sample of high-resolution SZE measurements is needed. A full calculation would also need to take into consideration effects such as detection apertures and the spatial filtering due to imaging algorithms, some of which would increase the importance of the effect and some of which would decrease its importance. This is the one of a very few clusters that has been observed at sub-arcminute resolution in the SZE \\citep[see also][]{nord09}, so it is possible that many clusters exhibit similar behavior. Such events, if their enhancement of the SZE brightness is transient, could also bias surveys towards detecting kinematically disturbed systems near the survey detection limits. The astrophysics that has been revealed by high resolution X-ray observations, and is beginning to be revealed by high resolution SZE data, is interesting in its own right. The SZE observations require large-aperture millimeter telescopes which have henceforth been lacking, but with both large single dishes and ALMA coming online, exciting observations will be forthcoming. There is substantial room for improvement: since the observations we report here the GBT surface has improved from $320 \\micron$ RMS to $250 \\micron$ RMS, which will yield more than a factor of $1.5$ improvement in sensitivity. The array used in these observations, while state of the art, has not yet achieved sky photon noise limited performance; further progress is being made in this direction. Considering these facts, and that the results presented here were acquired in a short period of allocated telescope time (8h), this new high-resolution probe of the ICM has a bright future. The National Radio Astronomy Observatory is a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc. We thank Eiichiro Komatsu for providing the NOBA SZ map; Marusa Bradac for providing her total mass map; Masao Sako, Ming Sun, Maxim Markevitch, Tony Mroczkowski and Erik Reese for helpful discussions; and Rachel Rosen and an anonymous referee for comments on the manuscript." }, "0910/0910.2433_arXiv.txt": { "abstract": "We have performed three searches for high-frequency signals in the solar neutrino flux measured by the Sudbury Neutrino Observatory (SNO), motivated by the possibility that solar $g$-mode oscillations could affect the production or propagation of solar $^8$B neutrinos. The first search looked for any significant peak in the frequency range 1/day to 144/day, with a sensitivity to sinusoidal signals with amplitudes of 12\\% or greater. The second search focused on regions in which $g$-mode signals have been claimed by experiments aboard the SoHO satellite, and was sensitive to signals with amplitudes of 10\\% or greater. The third search looked for extra power across the entire frequency band. No statistically significant signal was detected in any of the three searches. ", "introduction": "\\label{sec:intro} Neutrinos are the only way known to directly probe the dynamics of the solar core~\\citep{Bahcall:1988}, and, through the Mikheev-Smirnov-Wolfenstein (MSW) effect~\\citep{Mikheev:1986, Wolfenstein:1977}, they can even carry information about the rest of the solar envelope. To date, however, converting measurements of solar neutrino fluxes into constraints on solar models has proven to be difficult~\\citep{nussm}, because of the large number of co-varying parameters upon which such models are built. A relatively simple signal that could tell us something new about the Sun would be a time-variation in the neutrino fluxes. Over the past forty years, measurements made by solar neutrino experiments have therefore been the focus of many studies, ranging from attempted correlations with the solar sunspot cycle to open searches for signals with periods of weeks or months~\\citep{sturrock:03, sturrock:04, sturrock:05, sk:periodicity, sno:periodicity}. The shortest period examined to date is roughly one day, where the MSW effect predicts that neutrinos propagating through the Earth's core during the night will undergo flavor transformation in much the same way they do in the Sun, resulting in a net gain in the flux of electron neutrinos ($\\nu_e$s). Although there have been occasional claims of signals on timescales similar to known variations in the solar magnetic field, in all cases there have been conflicting measurements that show the signals to be spurious or absent entirely. We present in this article the results of a search in a new frequency regime for solar time variations. Our focus has been on signals whose periods range from 24 hours down to 10 minutes. The motivation for such a high frequency search is in part the expectation for solar helioseismological variations on scales of order an hour or less, in particular solar `gravity modes' ($g$-modes)~\\citep{christensen}. These $g$-modes are non-radial oscillations that are predicted to be confined to the solar core, and thus could in principle affect either neutrino production or neutrino propagation. The neutrinos that SNO detects, those from $^8$B decay within the Sun, are particularly well-suited for our search because they are created very deep within the solar core and because their propagation is known to be sensitive to variations in the solar density profile through the MSW effect. The effects of $g$-modes on solar neutrino fluxes have been examined by Bahcall and Kumar~\\citep{bk:gmode}, who sought to determine whether $g$-mode effects could explain the apparent solar neutrino deficit, finding that any effect was far too small to account for the roughly 60\\% discrepancy. More recently, ~\\citet{burgess} looked at ways in which a broad spectrum of $g$-modes could alter the expectation for a solar neutrino spectral distortion caused by the MSW effect. Nevertheless, there are at this time no explicit predictions as to whether $g$-modes or any other short-timescale variations could lead to measurable solar neutrino flux variations. ", "conclusions": "We have performed three searches for high-frequency signals in the $^8$B solar neutrino flux, applying a Rayleigh Power technique to data from the first two phases of the Sudbury Neutrino Observatory. Our first search looked for any significant peak in a Rayleigh Power spectrum from frequencies ranging from 1/day to 144/day. To account for SNO's deadtime window, we calculated the expected distribution of power in each bin of the Rayleigh Power spectrum using a random walk model, thus allowing us to assign confidence levels to the observed powers. We found no significant peaks in the data set. For this `open' peak search, we had a 90\\% probability of making a 99\\% CL detection of a signal with an amplitude of 12\\% or greater, relative to SNO's time-averaged neutrino flux. In a second search, we narrowed our frequency band to focus on a region in which $g$-mode signals have been claimed by experiments aboard the SoHO satellite. The examined frequency range extended from 18.5/day to 19.5/day. Again, no significant peaks in the Rayleigh Power spectrum were found, and our sensitivity for this `directed' search gave us a 90\\% probability of making a 99\\% CL detection for signals whose amplitudes were 10\\% or larger, relative to SNO's time-averaged neutrino flux. Our third search examined the entire range of frequencies from 1/day to 144/day, looking for any evidence that additional power was present across the entire high-frequency band. To do this, we used the distribution of frequency-specific confidence levels, determined using our random walk model. We found that, as expected for no high-frequency variations, this distribution was flat. We showed that for a simple Gaussian white noise model, the confidence level distribution would be noticeably distorted even when the amplitudes of the contributing frequencies have an rms as small as 0.1\\%." }, "0910/0910.5480_arXiv.txt": { "abstract": "We study the structure of neutron stars in $f(R)$ gravity theories with perturbative constraints. We derive the modified Tolman-Oppenheimer-Volkov equations and solve them for a polytropic equation of state. We investigate the resulting modifications to the masses and radii of neutron stars and show that observations of surface phenomena alone cannot break the degeneracy between altering the theory of gravity versus choosing a different equation of state of neutron-star matter. On the other hand, observations of neutron-star cooling, which depends on the density of matter at the stellar interior, can place significant constraints on the parameters of the theory. ", "introduction": "\\label{INTROsec} Recent interest in modified theories of gravity has been spurred by the discovery that the Universe is undergoing accelerated expansion (see, e.g.,~\\cite{P99,R98,WMAP09}). The simplest solution consistent with these observations posits a cosmological constant $\\Lambda$. The magnitude of this cosmological constant is significantly less than what was expected, and many undertakings have been made to see if there are plausible alternative explanations~\\cite{CDTT04,S08}. Outstanding questions also present themselves in the formation of singularities~\\cite{P08} and the seeming contradiction between quantum mechanics and gravity in the context of black hole thermodynamics~\\cite{W01}. All these suggest that there may yet be much to understand about the nature of gravity at extreme-curvature scales, far removed from our everyday experience. The two most popular approaches to modifying gravity have been the introduction of an additional scalar field (e.g.~\\cite{PR03}), or the related approach of replacing the Einstein-Hilbert action with a general function of the Ricci scalar $f(R)$ (e.g.~\\cite{CDTT04}). Within either framework the additional scalar degree of freedom can be tuned to mimic the cosmological constant, or any type of cosmological evolution at cosmological scales~\\cite{W07}. Despite the premise of such modifications, the non-linear character of gravitational theories has proven a significant obstacle to introducing new dynamical fields to drive modifications to gravity at the cosmological scale without the same fields reemerging at widely different curvature scales. One such example is the problem of ensuring that $f(R)=R\\pm\\mu^4/R$ theories pass the current Parametrized Post-Newtonian (PPN) bounds. When the new field is dynamical, the PPN parameter $\\gamma$ is forced to a value of $1/2$, which is very far from the present experimental bound~\\cite{C03}. As a result one has to choose a function $f(R)$ only from the class which can adequately suppress the new dynamical field on solar-system scales. The chameleon mechanism~\\cite{KW04,FTB07,HS08} provides such an alternative. In addition to the PPN constraints, instabilities related to the functional form of $f(R)$ have also been studied at length. This is especially true for the Dolgov-Kawasaki instability~\\cite{DK03}, which requires that $\\partial ^2 f/ \\partial R^2 > 0$ in order that the effective mass of the equivalent scalar degree of freedom be positive. In the strong-field regime, recent results~\\cite{KM08} suggest that this very choice may well prohibit the formation of compact objects above a curvature scale readily observed. However, the fatal curvature singularity may be avoided by the chameleon mechanism~\\cite{BL09,UH09}. Perhaps the source of the instabilities and consistency issues many of these models encounter is the result of treating these modifications as though they are exact. The original motivation behind introducing additional functions of the curvature was to generate a new phenomenology at a specific scale. However, many of the problems encountered by $f(R)$ gravity theories originate at curvature scales far removed from the ones under consideration. An alternative formulation for handling corrections to General Relativity is to view the new terms as only the next to leading order terms in a larger expansion. In this context there is no reason to suspect that the new phenomenology is due to new dynamical fields. The technique for handling a field expansion of this form is well developed~\\cite{EW89} and is known as perturbative constraints or order reduction~\\cite{JLM86}. Gravity with perturbative constraints allows us to explore alternative phenomenologies of gravity while maintaining important consistency conditions including gauge invariance, the assumption that we are approximating a fundamentally second order field theory, and the conservation of stress-energy. Maintaining such constraints while enlarging the space of possible behaviors of gravitation is the goal also of the Parametrized Post-Friedman approach~\\cite{HS07,H08,FS10}. In previous works~\\cite{DP08, CDP09}, we have analyzed the effect of treating $f(R)$ models of gravity via perturbative constraints primarily at cosmological scales. In this paper, we examine the ramifications of modifications to gravity in the context of compact objects. We show how the method of perturbative constraints allows for a consistent phenomenology for gravity on both large (Hubble-length perturbations linear in metric variables, but strongly relativistic, $L\\sim c/H_0$) and small scales (stellar scales, non-linear in metric perturbations, and strongly relativistic, $GM/rc^2\\sim 1$.) The layout of this work is as follows. In Section~\\ref{PCsec}, we review the equations of $f(R)$ gravity treated with perturbative constraints. In Section~\\ref{SPCsec}, we derive the modified Tolman-Oppenheimer-Volkov equations and show that the exterior solution is the Schwartzchild-de Sitter metric. In Section~\\ref{MRsec}, we demonstrate that such objects are stable and we solve numerically for their mass-radius relation for a polytropic equations of state. Finally in Section~\\ref{DIS} we discuss how we can discriminate modifications to gravity from uncertainty in the neutron star equation of state. ", "conclusions": "\\label{DIS} The predicted mass-radius relation for neutron stars in $f(R)$ gravity shown above differs from that computed within general relativity. However, very similar deviations in the mass-radius relation can also be obtained within general relativity by simply changing the polytropic index of the equation of state (see~\\cite{OP09} for examples). Because the equation of state of neutron-star matter is weakly constrained by current experiments, neutron-star observables that depend only on the mass and radius of the star cannot distinguish between small differences in the equation of state versus small modifications to gravity. In~\\cite{P08-2} it was shown that observables that depend also on the effective surface gravity of neutron stars can break, in principle, this degeneracy. In particular it was shown that the Eddington luminosity $L_E^{\\infty}$ of a bursting neutron star depends directly on its effective surface gravity as \\begin{equation} L_{\\rm E}^\\infty \\equiv \\frac{4\\pi m_{\\rm p} r_{\\rm s}} {(1+X)\\sigma_{\\rm T}} \\left[\\frac{z_{\\rm s}(z_{\\rm s}+2)}{(1+z_{\\rm s})^3}\\right] \\eta\\label{eq:obs2}\\;. \\end{equation} In this equation, $m_{\\rm p}$ is the mass of the proton, $X$ is the hydrogen mass fraction in the neutron-star atmosphere, $\\sigma_{\\rm T}$ is the Thomson scattering cross section, and \\begin{equation} z_{\\rm s} = \\left(1-\\frac{2M}{R}\\right)^{-1} -1 \\end{equation} is the gravitational redshift from the neutron star surface. The parameter $\\eta$ is the ratio of the effective surface gravity of the neutron star to that calculated in GR, i.e., \\begin{equation} \\eta\\equiv\\frac{g_{\\rm eff}}{g_{\\rm GR}} \\end{equation} with \\begin{equation} g_{\\rm eff}\\equiv \\left. \\frac{1}{2\\sqrt{A}}\\frac{d \\ln B}{dr}\\right|_{r=R} \\end{equation} and \\begin{equation} g_{\\rm GR}=\\frac{1}{2R}\\left[\\frac{z_s\\left(z_s+2\\right)}{z_s+1}\\right]\\;. \\end{equation} \\begin{figure}[t] \\includegraphics[angle=-90,scale=0.4]{alpha_rho.eps} \\caption{The central density $\\bar{\\rho}$ of neutron stars with different masses $\\bar{M}$ as a function of the parameter $\\bar{\\alpha}$. Larger positive deviations from general relativity require larger central densities for the same neutron-star mass and, therefore, lead to shorter cooling times. On the other hand, larger negative deviations require smaller central densities and lead to longer cooling time.} \\label{alpharhocgraph} \\end{figure} We can calculate easily the value of the parameter $\\eta$ for the $f(R) = R^2$ theory considered here. From the conservation equation~(\\ref{DT}) we can write \\begin{equation} \\label{geff} g_{eff}=-\\frac{1}{\\sqrt{A}}\\frac{P'}{\\left(\\rho+P\\right)}\\;. \\end{equation} We can then evaluate the hydrostatic equilibrium equation~(\\ref{HYDRO}) to first order in $\\alpha$ by noting that \\begin{equation} \\frac{R^{(0)'}}{A^{(0)}} = -8\\pi\\left(\\frac{\\partial \\rho_0}{\\partial P_0}-3\\right)\\frac{\\left(\\rho_0 + P_0\\right)}{r^2}\\left(M^{(0)}+4\\pi P_0 r^3\\right)\\;. \\end{equation} As a result equation~(\\ref{geff}) becomes \\begin{eqnarray} g_{eff} &=& \\frac{ \\sqrt{A^{(1)}} }{r^2} \\left(M^{(1)}+4\\pi P_1 r^3\\right) -\\alpha \\left\\{8\\pi \\left(\\rho_0+P_0\\right)\\sqrt{A^{(0)}} r \\right. \\nonumber \\\\ && \\left[ 2\\pi \\left(\\rho_0-3P_0\\right) + \\frac{2}{r^3}\\left(3-\\frac{\\partial \\rho_0}{\\partial P_0}\\right)\\left(M^{(0)}+4\\pi r^3 P_0\\right) \\right. \\nonumber \\\\ && \\left. \\left. \\left. +\\frac{A^{(0)}}{r^4}\\left(3-\\frac{\\partial \\rho_0}{\\partial P_0}\\right) \\left(M^{(0)} +4\\pi r^3 P_0\\right)^2 \\right]\\right\\}\\right|_{r=R} \\end{eqnarray} At the surface layer of the neutron star $\\rho = P = 0$ and hence \\begin{equation} g_{eff} = \\frac{\\sqrt{A^{(1)}}}{R^2}M^{(1)}\\;. \\end{equation} Which has the same dependence on mass and radius as $g_{GR}$ does. As a result measuring $\\eta$ alone will not suffice to break the degeneracy due to the equation of state. Nevertheless constraining observationally the cooling rates of neutron stars can offer a discriminant. A neutron star cools both through photon and neutrino emission. The photon luminosity is determined by the temperature at the photosphere, which in turn depends on the density of the photosphere. However neutrino cooling, which depends more sensitively on temperature than photon cooling does, becomes dominant for neutron stars with temperatures above $10^{10}K$, and indeed is the primary mechanism of cooling for young neutron stars (see~\\cite{PGW06} for a detailed review). The high temperature and low interaction rate make neutrino cooling particularly sensitive to the central density of the neutron star. Figure~\\ref{alpharhocgraph} shows the relation between the parameter $\\bar{\\alpha}$ and the central density of a neutron star, for three different values of the mass $\\bar{M}=0.15$, $0.125$, and $0.1$. Large positive deviations from general relativity, as measured by the parameter $\\bar{\\alpha}$ require larger central densities for neutron stars of a given mass, whereas the opposite is true for large negative deviations. As a result, because the cooling timescale scales with central density, observations of the surface temperatures of young neutron star can lead to useful constraints on the deviations from general relativity in an $f(R)$ gravity model, especially if the neutron-star masses are known. We will study the constraints imposed on $f(R)$ gravity by current measurements of cooling rates of neutron stars in our galaxy in a forthcoming paper." }, "0910/0910.5449_arXiv.txt": { "abstract": "The next generation of telescopes will acquire terabytes of image data on a nightly basis. Collectively, these large images will contain billions of interesting objects, which astronomers call \\textit{sources}. The astronomers' task is to construct a catalog detailing the coordinates and other properties of the sources. The source catalog is the primary data product for most telescopes and is an important input for testing new astrophysical theories, but to construct the catalog one must first detect the sources. Existing algorithms for catalog creation are effective at detecting sources, but do not have rigorous statistical error control. At the same time, there are several multiple testing procedures that provide rigorous error control, but they are not designed to detect sources that are aggregated over several pixels. In this paper, we propose a technique that does both, by providing rigorous statistical error control on the aggregate objects themselves rather than the pixels. We demonstrate the effectiveness of this approach on data from the Chandra X-ray Observatory Satellite. Our technique effectively controls the rate of false sources, yet still detects almost all of the sources detected by procedures that do not have such rigorous error control and have the advantage of additional data in the form of follow up observations, which will not be available for upcoming large telescopes. In fact, we even detect a new source that was missed by previous studies. The statistical methods developed in this paper can be extended to problems beyond Astronomy, as we will illustrate with an example from Neuroimaging.\\\\ \\textbf{Keywords:} Blind Source Detection, Multiple Testing, Astrostatistics ", "introduction": "The typical astronomical image records the intensity of light, over some range of frequencies, across a section of sky that contains many celestial objects of various size, shape, and luminosity. The image's pixels correspond to an array of light-sensitive detectors in the telescope, % and each pixel essentially counts how many photons have struck the corresponding detector during the exposure. But the photons recorded in the image do not come solely from the objects of interest, or \\emph{sources}; thermal noise and background emissions (collectively called \\emph{background}) corrupt the data and obscure the signature of the objects. Moreover, diffraction and atmospheric effects blur the image, reducing resolution and washing out the fainter signals. In the (astronomical) source detection problem, one is given such an image and seeks to construct a \\emph{catalog} that gives the coordinates (and often other properties) of sources in the image. A source catalog is the basic data product of most astronomical surveys and the basic input to the scientific process. This has been true for some time. Early catalogs -- from the data of ancient astronomers Shi Shen and Hipparchus, each cataloging about 1000 stars, to the compendium of deep-sky objects produced by William and Caroline Herschel in the 1700s \\citep{catalogue-of-nebulae} % -- were based on direct visual observations. Later work, especially in the 20th century, used photographic plates, both improving resolution and allowing the detection of much fainter objects. But either way, compiling a source catalog would be a slow and painstaking affair, often requiring years to collect data on only a handful of objects. Until recently, catalogs comprising a few hundred objects were large, a few thousand were epic. All this changed with the advent of new technologies -- digital imaging, advanced designs for telescope mirrors, and computer automation -- and with increases in available computing power and storage. With relative suddenness, astronomers found that they could observe wider, deeper, and faster than ever before. They could sweep the sky searching automatically for objects of a certain type, they could collect data on many objects in parallel, and they could observe a multitude of faint objects that would previously have gone undetected. The Sloan Digital Sky Survey \\citep{sdss} has measured hundreds of millions of objects. The upcoming Large Synoptic Survey Telescope (LSST, \\citep{lsst}) will scan the entire sky every few days, collecting several terabytes of data per night into a catalog comprising \\emph{billions} of objects. Over the past two decades, astronomy has gone from data poor to data rich. And therein lies both opportunity and challenge. The opportunity lies in the richness and importance of the scientific questions that these massive data sets can answer. The challenge lies in the sheer scale of the data analysis. While as a general rule in science, more data is better, there is a reason that astronomers often describe the coming bounty of data in quasi-biblical terms -- a flood, a tsunami, an onslaught. The next generation of astronomical catalogs, including the LSST, will be so large that even simple operations -- such as a basic query of the entire catalog -- will be computationally prohibitive, and yes that does account for Moore's law. This influx of data has motivated several statistical innovations in the field of Astronomy leading to the emergence of the sub-field, astrostatistics. New statistical innovations have had a significant impact on several important and cutting edge Astronomy problems, for a small sampling see \\citet{vandyk-2009-3}, \\citet{loredo}, \\citet{rice}, \\citet{npicmb}, and \\citet{richards}. Besides the massive size of the data, another issue is controlling the rate of errors in the catalog. In the past, where every object in the catalog was observed manually, sources were often missed, located incorrectly, or created spuriously. Follow-up observations can often be made for all or most of the sources, reducing false positive and false negative identifications to a manageable level. But in the near future, the sheer number of objects in the catalog will preclude comprehensive follow-up observations by human beings. Scientific studies of the catalog will likely need to be based on samples or selections made by automatic, statistical criteria. But no matter how carefully these criteria are constructed, there will be objects that are misclassified. To use the resulting samples effectively for scientific inference, it will be necessary for the method to provide tunable control of error rates. Thus, new statistical and computational methods will be needed to construct and analyze the next generation of astronomical source catalogs. In this paper, we develop a multiple-testing-based method for the source detection problem that has several advantages over existing techniques, especially for the analysis of large-scale surveys like the LSST. Although we discuss our method in the context of astronomical source detection, the method applies to a wide range of similar problems such as neuroimaging \\citep[e.g.][]{neuro} and remote sensing \\citep[e.g.][]{remote}, and we give such an example in a later section. We assume that the input image is an $n\\times m$ array of pixels, with the value recorded at pixel $(i,j)$ denoted by $Y_{ij}$. The photons that contribute to $Y_{ij}$ arise from two components: \\emph{sources}, the emissions produced by the celestial objects of interest, and \\emph{background}, which includes thermal noise, the emissions of unresolved objects, interfering radiation sources, atmospheric emissions, and all other anomalies or artifacts. astronomical images essentially measure photon counts for which a Poisson model is appropriate \\citep{poissModel}. So for our base model, we assume that the $Y_{ij}$'s are independent with \\begin{equation} \\label{eq::simple-poisson-model} Y_{ij} \\;{\\rm distributed\\ as\\ } {\\rm Poisson}\\langle \\lambda_{1,ij} + \\lambda_{0,ij}\\rangle, \\end{equation} where $\\lambda_{1,ij} \\ge 0$, $\\lambda_{0,ij} \\ge 0$ denote the mean intensity of sources and background, respectively and $\\lambda_{1,ij}+\\lambda_{0,ij} > 0$. The idea here is that the pixels are measuring the counts in disjoint cells of a Poisson random field across the sky. This applies to good approximation for space-based observations like those reported in Section 2. For ground-based observation, the image is in addition blurred by atmospheric turbulence, so the $Y_{ij}$'s are no longer strictly independent. However, the Poisson model in Equation \\ref{eq::simple-poisson-model} still holds to reasonable approximation. In data sets where the counts are high, the Poisson random field can be further approximated by a Gaussian random field. In this paper, we utilize a technique originally developed for the Gaussian model and generalize it so that we can accommodate a wide variety of models. The source detection problem is to identify which pixels contains sources and thus to separate the sources from the background. If $\\lambda_{1,ij} > 0$, then we take pixel $(i,j)$ to be a source pixel; otherwise, it is a background pixel. So it is natural to consider this as a multiple testing problem with the null hypothesis at each pixel being that $\\lambda_{1,ij} = 0$. At a coarse level, we want to characterize the set $\\cS = \\Set{(i,j):\\; \\lambda_{1,ij} > 0}$ of source pixels, but our more specific goal is to identify and locate the underlying sources, so that an accurate catalog can be constructed. This requires a more stringent criterion for success because the objects are coherent, localized aggregates. As Figure \\ref{fig:type1and2} shows, with the same number of pixel-wise type I and type II errors, it is possible to get widely varying accuracy in the resulting catalog. Put another way, our loss function operates on the catalog, not the pixels themselves. \\begin{figure}[h!] \\centerline{\\includegraphics[scale=.35]{typeIandII.ps}} \\caption{The 45 pixels that have any overlap with the red circle are considered sources and black pixels indicate sources detected via some detection algorithm. Each of the three images have 6 Type I error pixels and 24 Type II error pixels. The detection in the left image captures the center of the source but clearly misses the shape. The detections in the center image show several different sources that have some overlap with the true source. The detection in the right image captures the center of the true source and has some spurious noise detections. While all these pictures have the same number of pixel-wise Type I and Type II errors they lead to very different conclusions about the number, shape and location of the sources in the image, thus a testing criteria based on the aggregate sources, instead of pixels, is necessary. } \\label{fig:type1and2} \\end{figure} In this paper, we extend the False Cluster Proportion (FCP) controlling procedures introduced by Perone Pacifico et al. (2004) to make it effective for controlling the rate of false sources detected in astronomical images. As we will show below, the original FCP procedure does not perform well with the Poisson statistics common in the source detection problem and even where the Gaussian assumption holds, does not yield sufficient power to be viable for the astronomical source detection problem. We generalize the technique so that it applies to a wider range of noise models. We also introduce a new transform, which we call the \\emph{Multi-scale Derivative}, that enhances sources and significantly improves power. Taken together, these extensions lead to a new procedure that we call the Multi-scale False Cluster Proportion (MSFCP) procedure. This gives a powerful source detection technique that provides rigorous control over the rate of false \\emph{sources}, where techniques in current use provide control over the rate of false \\emph{pixels}, if they provide any control at all. We demonstrate both the excellent power and error control on a very deep and high resolution telescope image from the Chandra X-Ray Observatory. MSFCP has detection power competitive with the existing procedures, even detecting a source that was overlooked in previous studies while at the same time maintaining rigorous control over the rate of false sources. Although the source detection problem is ubiquitous in Astronomy it also occurs in other settings and we present an example of how FCP concepts can be extended to detecting bands of neural activity in the brain. Due to its prevalence and difficulty, there have been a variety of approaches to the source detection problem, both in the statistical and astronomical literature. Given the recent explosion in research on multiple testing, there is a plethora of available techniques that can be applied directly to the pixel-wise hypothesis tests to reconstruct $\\cS$, including \\citet{bh}[BH], \\citet{stepup}, \\citet{storey}, \\citet{suncai}, and \\citet{rice} . However, because these methods give error-rates in terms of the individual pixel-wise tests, it is not obvious how to translate these error-rates to make inferences about the underlying sources. Multiple testing approaches that are less pixel-centered have been developed in the related problem of analyzing functional magnetic resonance imaging (fMRI) data. In this problem, the sources are regions of neural activity that reveal themselves through a measurable change in blood flow response. \\cite{wor96} and \\citet{wor02} use level sets of a random field formed from test statistics to identify regions containing sources. \\citet{cba} construct a test based on clusters instead of pixels in the fMRI setting, but this technique takes advantage of a temporal dimension that is not available in many general (e.g., astronomical) problems. Source detection understandably garners much attention in the astronomical literature. Astronomers address source detection as one step in a data-processing pipeline -- the series of operations performed on the data from collection until catalog. These include, but are not limited to, corrections for atmospheric effects, image registration, filtering out unwanted signals, as well as source detection. These pipelines are typically planned and developed well before the instrument is operational. During the planning stage, simulations are run to test the pipeline including the source detection algorithm \\citep[e.g.,][]{sehgal}. Astronomers typically do not require their algorithms to satisfy any formal performance criteria; rather, they apply an empirical criterion, using data simulated to look like a real telescope image to calibrate the error rate for detected sources that will be expected in practice. Of course, this depends on the simulated and real data being both quantitatively and qualitatively similar. While great effort and ingenuity are applied in constructing realistic simulations -- sometimes years of computing time for a single run -- the simulations still rely on untested and unstated assumptions that may fail when the instrument comes on line. The source detection methods used by astronomers fall into three general classes: simple thresholding, peak-finding algorithms, and Bayesian algorithms. Simple thresholding consists of choosing an intensity cutoff for pixel-wise statistics and classifying any pixel above threshold as a source. It is popular because it simple and fast, easily computed by the popular SExtractor software \\citep{sext}. Before thresholding, filters are often applied to the raw data to suppress confounding background signals. In \\citet{vik} and \\citet{melin}, Matched Filters are used to isolate the signal, and then thresholds are determined using simulated data to create catalogs in X-ray and Radio telescope images respectively. \\citet{2002AJ....123.1086H} also use simple thresholding, choosing the threshold to control the False Discovery Rate via the method of \\citet{bh}, but they use simulated telescope images to calibrate between rate of false pixels and the rate of false sources. Peak-finding algorithms search for local maxima in the denoised image and catalog them as sources. \\citet{vale} and \\citet{sehgal} use peak-finding algorithms to look for large galaxy clusters in simulated radio telescope images. The wavelet-based technique of \\citet{peter} is commonly used to detect X-ray sources as in \\citet{valt} and \\citet{1msec}. \\citet{Damiani} and \\citet{gonzaleznuevo-2006-369} provide a good overview of the popular Mexican Hat wavelet as a tool for source detection. Several implementations exist in software and most are specifically tailored for a certain instrument or type of problem. Pixel-wise error rates for wavelet source detection can sometimes be determined analytically but are more often approximated from simulations. As with simple thresholding, simulations are often used to indirectly estimate the rate of false sources, the error rate of interest, from the rate of false pixels. Bayesian techniques have become popular with astronomers and several have applied Bayesian methods to source detection as in \\citet{2007ApJ...661.1339S}, \\citet{strong}, and \\citet{2003MNRAS.338..765H}. Bayesian detection algorithms typically define models for sources, often two-dimensional Gaussians, and attempt to distinguish them from backgrounds via inference on a posterior distribution. \\citet{gug} use Bayesian mixture models to separate sources from background, while \\citet{fastBayes} propose ways to speed up Bayesian source detection, which can be computationally slow for large problems. Bayesian algorithms typically require more assumptions to be made a priori and can be more computationally expensive than thresholding or peak-finding algorithms. Our goal in source detection is to detect the relevant objects, not pixels, while controlling the error rates. Statisticians have developed numerous methods for dealing with error rates but have not been focused on detecting aggregate objects, while astronomers have been thinking about detecting objects, but without a rigorous approach to controlling error rates. We propose a technique that does both: control the error rates and make our inference about the sources themselves. We demonstrate our techniques on an important data set from the Chandra X-ray Observatory, one of the most powerful telescopes in existence. For the Chandra data, our goal is to detect the X-ray sources with a bound on the error rate for sources. We describe an approach due to \\citet{Pacifico04} that gives us a probabilistic bound on the error rate for sources, and apply it to the Chandra data. We find that while this approach gives us the error control we want, it does not have good power when compared to other techniques. We introduce a generalization of the technique that allows for us to keep a probabilistic bound on the error rate for sources under more general conditions. We then introduce a new Multi-scale technique that is designed to enhance sources and thus increase power. We then integrate it with our generalized procedure to get the MSFCP procedure which increases power while maintaining control over the error rate. This improvement is evident when we revisit the Chandra data -- we show that our power using MSFCP is competitive to two algorithms commonly used by astronomers, but with superior error control. Furthermore, using our procedure we detect a X-ray source which had gone undetected in the original analysis of the data by astronomers. We then provide a brief description of how these techniques can be used outside the realm of Astronomy with an application to high-resolution neuroimaging data. We conclude with an overview of our results and directions for further study. ", "conclusions": "We have extended the False Cluster Proportion Multiple Testing Procedure so it can now be applied to a wide variety of problem in Astronomy as well as other fields. By moving away from theoretical approximations, like Piterbarg's approximation, we can now calculate confidence supersets for a large variety of noise conditions. This allows us to create source catalogs with a probabilistic guarantee that they are not overly polluted with false detections without having to make any difficult to check assumptions about the data or the science behind the data. The Multi-scale Derivative procedure enhances the types of sources we typically see in a telescope image and we have shown that they can also enhance the power of False Cluster Proportion source detection algorithms. This is evidenced by our analysis of the CDFS data. The catalog published in GI uses two different algorithms and then has to follow up each detection, whereas we run MSFCP and can make the statement that with probability .95 less than 10\\% of our detections are false, and still get virtually the same catalog. When we allow 20\\% of our detections to be false we make two new detections, one of which we have verified is a real source that was missed in the original analysis. Our procedure provides rigorous error control, which is lacking in the current techniques used by astronomers. At the same time we have demonstrated that the detection power is comparable and in fact we have detected sources that they have missed. Controlling false detections without the need for expensive follow up observations will be critical as the next generation of telescopes will provide a deluge of data that will be impossible to process manually. We have also shown that these ideas can be generalized to other types of problems where we are dealing with points spread along a plane instead of a regular grid of pixels in an image. We can further modify our definition of cluster to differentiate between different types of objects. We can then control the proportion of false clusters, allowing for different classes of clusters. In the future, we believe we can further extend our framework so that we can apply False Cluster Proportion techniques on a wide variety of statistical clustering problems, in which we are trying to detect clusters of data without making false detections. {\\scriptsize" }, "0910/0910.3954_arXiv.txt": { "abstract": "We study the dynamics of stellar-mass black holes (BH) in star clusters with particular attention to the formation of BH-BH binaries, which are interesting as sources of gravitational waves (GW). In the present study, we examine the properties of these BH-BH binaries through direct N-body simulations of star clusters using the NBODY6 code on Graphical Processing Unit (GPU) platforms. We perform simulations for star clusters with $\\leq 10^5$ low-mass stars starting from Plummer models with an initial population of BHs, varying the cluster-mass and BH-retention fraction. Additionally, we do several calculations of star clusters confined within a reflective boundary mimicking only the core of a massive star cluster which can be performed much faster than the corresponding full cluster integration. We find that stellar-mass BHs with masses $\\sim 10\\Ms$ segregate rapidly ($\\sim 100$ Myr timescale) into the cluster core and form a dense sub-cluster of BHs within typically $0.2 - 0.5$ pc radius. In such a sub-cluster, BH-BH binaries can be formed through 3-body encounters, the rate of which can become substantial in dense enough BH-cores. While most BH binaries are finally ejected from the cluster by recoils received during super-elastic encounters with the single BHs, few of them harden sufficiently so that they can merge via GW emission within the cluster. We find that for clusters with $N \\ga 5\\times 10^4$, typically 1 - 2 BH-BH mergers occur per cluster within the first $\\sim 4$ Gyr of cluster evolution. Also for each of these clusters, there are a few escaping BH binaries that can merge within a Hubble time, most of the merger times being within a few Gyr. These results indicate that intermediate-age massive clusters constitute the most important class of candidates for producing dynamical BH-BH mergers. Old globular clusters cannot contribute significantly to the present-day BH-BH merger rate since most of the mergers from them would have occurred much earlier. On the other hand, young massive clusters with ages less that 50 Myr are too young to produce significant number of BH-BH mergers. We finally discuss the detection rate of BH-BH inspirals by the ``LIGO'' and ``Advanced LIGO'' GW detectors. Our results indicate that dynamical BH-BH binaries constitute the dominant channel for BH-BH merger detection. ", "introduction": "Star clusters, \\eg, globular clusters (henceforth GC), young and intermediate-age massive clusters and open clusters harbor a large overdensity of compact stellar remnants compared to that in the field by virtue of their high density and the mass segregation of the compact remnants towards the cluster core \\citep{hv83}. These compact stars, which are neutron stars (henceforth NS) and black holes (henceforth BH), are produced by the stellar evolution of the most massive stars within the first $\\sim$ 50 Myr after cluster formation. These compact stars, being generally heavier than the remaining low-mass stars of the cluster, segregate quickly (within one or a few half-mass relaxation times) to the cluster core, forming a dense sub-cluster of compact stars. Such a compact-star sub-cluster is of broad interest as it efficiently produces compact-star binaries through dynamical encounters \\citep{hv83,h.et.al92}, \\eg, X-ray binaries and NS-NS and BH-BH binaries, which are of interest for a wide range of physical phenomena. For example, X-ray binaries are the primary sources of GC X-ray flux, while mergers of tight NS-NS binaries (through emission of gravitational waves) is the most likely scenario for the production of short duration GRBs and both NS-NS and BH-BH mergers are very important sources of gravitational waves (henceforth GW) detectable by future gravitational wave observatories like ``Advanced LIGO'' (AdLIGO) and ``LISA'' \\citep{as2007}. Double-NS systems are also interesting because they could be observable as double-pulsar systems \\citep{rsm2008}, allowing important tests of General Relativity \\citep{krm2008}. In the present work, we investigate the dynamics of stellar-mass BHs in star clusters, with particular emphasis on dynamically formed BH-BH binaries. Such binaries are strong sources of GWs as they spiral-in through GW radiation, a process detectable out to several thousand Mpc distances. Tight BH binaries that can merge within a Hubble time can also be produced in the galactic field from tight stellar binaries, which result from stellar evolution of the components (\\eg, \\citealt{bul2003,osh2007,osh2008,bky2007}). However, \\citet{bky2007} have shown with revised binary-evolution models that the majority of potential BH-BH binary progenitors actually merge because of common envelope (CE) evolution which occurs when any of the binary members crosses the Hertzsprung gap. They found that this reduces the merger rate by a factor as large as $\\sim 500$ and the resulting AdLIGO detection rate of BH-BH binary mergers in the Universe from primordial binaries is only $\\sim 2$ yr$^{-1}$, subject to the uncertainties of the CE evolution model that has been incorporated. Note that the corresponding NS-NS merger rate, as the above authors predict, is considerably higher, $\\sim 20$ yr$^{-1}$. In view of the above result, the majority of merging BH binaries are possibly those that are formed dynamically in star clusters. As studied by several authors earlier \\citep{mt2006,mak2007}, black holes, formed through stellar evolution, segregate into the cluster core within $\\sim 0.3$ pc and form a sub-cluster of BHs, where the density of BHs is large enough that BH-BH binary formation through 3-body encounters becomes important \\citep{hh2003}. These dynamically formed BH binaries then ``harden'' through repeated super-elastic encounters with the surrounding BHs \\citep{h75,bg2006}. The binding energy of the BH binaries released is carried away by the single and binary BHs involved in the encounters. This causes the BHs and the BH binaries to get ejected from the BH-core to larger radii of the cluster and they heat the cluster while sinking back to the core through dynamical friction \\citep{mak2007}. Most of the energy of the sinking BHs is deposited in the cluster core as the stellar density is much higher there. As the BH binaries harden, the encounter-driven recoil becomes stronger and finally the recoil is large enough that the encountering single BH and/or the BH binary escape from the cluster (see Sec.~\\ref{res}, also \\citealt{bcq2002}). Because of the associated mass-loss from the cluster core, this also results in cluster heating. These heating mechanisms result in an expansion of the cluster, as studied in detail by several authors, \\eg, \\citet{mt2006,mak2007}. \\citet{mak2007} found notable agreement between the core expansion as obtained from their N-body simulations with the observed age-core radius correlation for star clusters in the Magellanic Clouds. The dynamics of stellar mass BHs and formation of BH binaries in star clusters and the resulting rate of GW emission from close enough BH binaries has been studied by several authors using N-body integrations \\citep{pzm2000}, numerical 3-body scattering experiments \\citep{gul2004} or Monte-Carlo methods \\citep{olr2006}. \\citet{pzm2000} considered the rate of mergers of escaping BH binaries from various stellar systems, \\eg, massive GCs, young populous clusters and galactic nuclei. They estimated the merger rates within 15 Gyr (their adopted age of the Universe) by assuming the binding-energy ($E_b$) distribution of the escaping BH binaries to be uniformly distributed in $\\log{E_b}$, as inferred from simulations of $N \\approx 2000$ or 4000 star clusters (see \\citealt{pzm2000} for details). Considering the space-densities of the different kinds of star clusters, they found that while for LIGO the corresponding total (\\ie, contribution from all types of star clusters) detection rate is negligible, it can be as high as $\\sim 1$ day$^{-1}$ for AdLIGO. \\citet{gul2004} performed sequential numerical integrations of BH binary-single BH close encounters in a uniform stellar background, where, in between successive encounters, the BH-binaries were evolved due to GW emission. From such simulations, these authors studied the growth of BHs through successive BH-binary mergers for the first time. In a more self-consistent study of the growth of BHs and the possibility of formation of intermediate mass black holes through successive BH-binary mergers, \\citet{olr2006} used the Monte-Carlo approach and considered ``pure'' BH clusters that are dynamically detached from their parent clusters which can be expected to form due to mass stratification instability \\citep{spz}. These authors considered dynamically formed BH binaries from 3-body encounters in such BH clusters utilizing theoretical cross-sections of 3-body binary formation (see \\citealt{olr2006} and references therein). Furthermore, they included both binary-single and binary-binary encounters. Considering the sub-set of BH binaries formed in the BH clusters that merge within a Hubble time, they determined typically a few AdLIGO detections per year for old GCs. While such results are remarkable and promising, a more detailed study of the dynamics of stellar-mass BHs in star clusters with a realistic number of stars is essential. As the cross-sections of different processes governing the dynamics of BH binaries, \\viz, 3-body binary formation, binary-single star encounters and ionization have different dependencies on the number of stars $N$ of the cluster, an extrapolation to much different $N$ can be problematic. Hence it is important to consider clusters with values of $N$ appropriate for massive clusters or GCs. Also, exact treatments of the various dynamical processes are crucial for realistic predictions of BH-BH merger rates. Finally, a more careful study of the different dynamical processes leading to the formation and evolution of BH binaries in star clusters is needed to understand better under which conditions tight inspiralling BH binaries can be formed dynamically from a star cluster. In the present work, we make a detailed study of the dynamics of BH-BH binaries formed in a BH sub-cluster, as introduced above. In particular, we investigate whether hard enough BH binaries that can merge via gravitational radiation in a Hubble time within the cluster or after getting ejected from the cluster, can be formed in such a sub-cluster. To that end, we perform direct N-body integrations of concentrated star clusters (half-mass radius $r_h \\leq 1$ pc) consisting of $N \\leq 10^5$ low-mass stars in which a certain number of stellar-mass BHs is added, representing a star cluster with an evolved stellar population. The present paper is organized as follows: In Sec.~\\ref{sim} we describe our simulations in detail. We discuss the various elements and assumptions of the simulations and summarize all the runs that we perform (Sec.~\\ref{runs}). We also discuss the use of a reflective boundary to simulate only the core of a star cluster (Sec.~\\ref{reflct}). In Sec.~\\ref{res} we discuss our results with particular emphasis on the dynamical BH binaries formed during the simulations. We discuss the BH-BH mergers that occur within the clusters and also the merger timescales of the BH binaries that escape from the clusters (Sec.~\\ref{mrgesc}) and obtain the distributions of merger-times for both cases (Sec.~\\ref{tdist}). Finally, in Sec.~\\ref{discuss} we interpret our results in the context of different types of star clusters that are observed and provide estimates of BH-BH merger detection rates. ", "conclusions": "\\label{discuss} Our present study indicates that centrally concentrated star clusters, with $N \\ga 4.5\\times 10^4$ are capable of dynamically producing BH binaries that can merge within a few Gyr, provided a significant number of BHs are retained in the clusters after their birth. The results of our simulations (see Table~\\ref{tab1}) imply that most of the BH-BH mergers occur within the first few Gyr of cluster evolution for both mergers within the cluster and mergers of escaped BH binaries. \\begin{figure} \\includegraphics[width=8.5cm,angle=0]{f5.eps} \\includegraphics[width=8.5cm,angle=0]{f6.eps} \\caption{{\\bf Top:} Distribution of the merger times $t_{mrg}$ for BH binary mergers within the cluster for the models of Table~\\ref{tab1}. {\\bf Bottom:} Distribution of the merger times $t_{mrg}$ for escaped BH binaries for the models of Table~\\ref{tab1} (see text).} \\label{fig:mrgdist} \\end{figure} The above results imply that an important class of candidates for dynamically forming BH binaries that merge at the present epoch are star clusters with initial mass $M_{cl} \\ga 3\\times 10^4 \\Ms$\\footnote{To correlate with the observed clusters, we consider the mass $M_{cl}$ of the parent cluster, \\ie, the cluster mass before the BHs are formed through stellar evolution, which are heavier than the clusters of low mass stars that we model, for the same value of $N$. For a given simulated cluster, we estimate the corresponding parent mass $M_{cl}$ by weighting the mean stellar mass of $\\langle m \\rangle_{cl}\\approx 0.6 \\Ms$ of a Kroupa IMF with star from $0.1\\Ms$ upto $100\\Ms$ with the total number of stars $N$ for that cluster.}, which are less than few Gyr old. Such clusters represent intermediate-age massive clusters (hereafter IMC) with initial masses close to the upper-limit of the initial cluster mass function (ICMF) in spiral \\citep{wkl2004,lrs2009a} and starburst \\citep{gls2006} galaxies. While it is not impossible to obtain BH-BH mergers within a Hubble time from lower-mass clusters, the overall BH-BH merger and escape rates strongly decrease with cluster-mass as Table~\\ref{tab1} indicates. For $M_{cl} \\la 1.5 \\times 10^4 \\Ms$, mergers already become much rarer (see Table~\\ref{tab1}). Due to the statistical nature of merger or ejection events it is ambiguous to set any well-defined limit on the cluster-mass beyond which these events become appreciable (such an estimate would also require a much larger number of N-body integrations). In view of our results, $M_{cl} \\approx 3\\times 10^4 \\Ms$ is a representative lower limit beyond which an appreciable number of mergers and escapers merging within a Hubble time can be obtained. Old globular clusters, which can be about 10 times or more massive, are expected to produce mergers or escapers more efficiently. As the timescale of depletion of BHs from the BH cluster is nearly independent of the parent cluster mass (see Sec.~\\ref{tdist}), GCs can also be expected to produce BH-BH mergers over similar time-span as the IMCs, \\ie, within the first few Gyr of evolution. Since GCs are typically much older ($\\sim 10 {\\rm ~Gyr}$), they do not contribute significantly to the present-day merger rate, since most of the mergers from them would have occurred earlier. Considering the light-travel time of $\\approx 4.5 {\\rm ~Gyr}$ from the maximum distance $D\\approx 1500 {\\rm ~Mpc}$ form which these BH-BH binaries can be detected by ``AdLIGO''(see below), only GCs close to the above distance could contribute detectable events, mostly from escaped BH-BH binaries. On the other hand, young massive clusters with ages less than 50 Myr, representing star clusters near the high-mass end of the ICMF \\citep{lrs2009b}, are generally too young to produce BH-BH mergers. All models in Table~\\ref{tab1} produce mergers significantly later than this age (except one escaped BH binary in each of the models C65K110 and C100K200). Hence, IMCs seem to be most likely candidates for producing observable BH-BH mergers dynamically. \\subsection{Detection rate}\\label{rate} We now make an estimate of the BH-BH merger detection rate from IMCs by ground-based GW observatories like LIGO and AdLIGO. In estimating the overall BH-BH merger rate using the results of our model clusters, one needs to consider the distribution of the cluster parameters that are varied over the models, \\viz, cluster mass, half-mass radius, and BH retention fraction. Such distributions are far from being well determined, except for the mass distribution for young clusters in spiral and starburst galaxies \\citep{bik2003,blm2003,gls2004}. Therefore, determination of an overall merger rate considering the distribution of our computed clusters can be ambiguous. Hence, as a useful alternative, we determine the BH binary merger detection rates for each of the cluster models in Table~\\ref{tab1} that gives an appreciable number of mergers, for a representative density of IMCs. Such an approach has been considered by earlier authors, \\eg, \\citet{olr2006} and can provide a reasonable idea of the rate of detection of BH-BH mergers from IMCs. As an estimate of the space density of IMCs, we adopt that for young populous clusters in \\citet{pzm2000}, which has a similar mass-range as the IMCs: \\begin{equation} \\rho_{cl} = 3.5 {\\rm ~} h^3 {\\rm ~Mpc}^{-3}, \\label{eq:rypc} \\end{equation} where $h$ is the Hubble parameter, defined as $H_0/100{\\rm ~km~s}^{-1}$, $H_0$ being the Hubble constant \\citep{pbl93}. The above space density has been derived from the space densities of spiral, blue elliptical and starburst galaxies \\citep{hey97} assuming that young populous clusters have the same specific frequencies ($S_{N}$) as old GCs \\citep{vdb95,mcl99}, but in absence of any firm determination of the $S_{N}$s of the former. We compute the detection rate for each model cluster assuming that it has a space density of the above value. The LIGO/AdLIGO detection rate of BH-BH mergers from a particular model cluster can be estimated from (\\citealt{bky2007} and references therein) \\begin{equation} {\\mathcal R}_{LIGO} = \\frac{4}{3}\\pi D^3\\rho_{cl}{\\mathcal R}_{mrg}, \\label{eq:ligorate} \\end{equation} where ${\\mathcal R}_{mrg}$ is the compact binary merger rate from a cluster and $D$ is the maximum distance from which the emitted GW from a compact-binary inspiral can be detected. $D$ is given by \\begin{equation} D = D_0 \\left(\\frac{M_{ch}}{M_{ch,nsns}}\\right)^{5/6}, \\label{eq:range} \\end{equation} where $D_0=18.4$ and 300 Mpc for LIGO and AdLIGO respectively. The quantity $M_{ch}$ is the ``chirp mass'' of the compact binary with component masses $m_1$ and $m_2$, which is given by \\begin{equation} M_{ch} = \\frac{(m_1 m_2)^{3/5}}{(m_1 + m_2)^{1/5}}, \\label{eq:chirp} \\end{equation} and $M_{ch,nsns}=1.2\\Ms$ is that for a binary with two $1.4\\Ms$ neutron stars. For a BH binary with $m_1=m_2=10\\Ms$, $M_{ch}=8.71\\Ms$ which gives $D \\approx 1500{\\rm ~Mpc}$ for AdLIGO. The AdLIGO detection rates ${\\mathcal R}_{\\rm AdLIGO}$ (mean over 3 Gyr taking into account the time of escape $t_{esc}$ of the escaped binaries) for the model clusters are shown in Table~\\ref{tab1}, where the currently accepted value of the Hubble parameter $h=0.73$ is assumed. The error in each detection rate is simply obtained from the Poisson dispersion of the total number of mergers for the corresponding cluster. Note that these detection rates are for clusters with solar-like metallicity which is implied by our assumption of $10\\Ms$ BHs for all the clusters. To obtain a basic estimate of the overall detection rate of BH-BH mergers from IMCs, we consider the subset of our computed models that are isolated clusters with full BH retention (see Table~\\ref{tab1}). We take the mass function of the IMCs to be a power law with index $\\alpha = -2$ which is the approximate index of the ICMF in spiral and starburst galaxies (see \\eg, \\citealt{gls2004,lrs2009a}). Then the weighted average of the corresponding AdLIGO detection rates is $\\overline{\\mathcal R}_{\\rm AdLIGO} \\approx 31(\\pm 7)$ yr$^{-1}$, which estimates the total present-day detection rate of BH-BH mergers from IMCs expected for AdLIGO. The corresponding LIGO detection rate is negligible, $\\overline{\\mathcal R}_{\\rm LIGO} \\approx 7.4\\times 10^{-3}{\\rm ~yr}^{-1}$. Note that these BH-BH detection rates are only lower limits. First, the observed population of star clusters can be an underestimation by a factor of 2 owing to their dissolution in the tidal field of their host galaxies (see \\citealt{pzm2000} and references therein). Second, the above detection rates are only from IMCs and there can be additional contributions from GCs (see above). In comparing the AdLIGO detection rates from our computations with those from earlier works, we note that our rates are typically an order of magnitude smaller than those of \\citet{pzm2000}, but about one order of magnitude larger than those obtained by \\citet{olr2006}. The principal origin of the former difference is due to the fact that while we considered only IMCs distinguishing them as the most appropriate candidates for producing present-day BH-BH mergers (see above), \\citet{pzm2000} also included GCs, which have considerably larger spatial density and also larger fraction of BH-BH binaries merging within the Hubble time, as obtained from their analytic extrapolations. Note that the number of escapers as obtained by them for young populous clusters and the fraction of them merging within a Hubble time (see Table~1 of \\citealt{pzm2000}) is similar to that obtained from the present computations, implying qualitative agreement. The above authors apparently did not consider the time scales of formation and depletion of the BH sub-systems in their preliminary study. On the other hand, although \\citet{olr2006} considered clusters significantly more massive than our's in their Monte-Carlo approach, so that larger merger detection rates can be expected, their clusters were much older (8 Gyr and 13 Gyr) than IMCs which, in accordance with our results, accounts for their much lower detection rate. It is interesting to note that the dynamical BH-BH merger detection rates obtained by us are typically an order of magnitude higher than that from primordial stellar binaries as predicted by \\citet{bky2007} based on their revised binary evolution model, and is similar to that for the isolated NS-NS binaries derived by them. Hence our results imply that dynamical BH-BH binaries constitute the dominant contribution to the BH-BH merger detection. Thus, the dynamical BH-BH inspirals from star clusters seem to be a promising channel for GW detection by the future AdLIGO, although their estimated detection rate with the present LIGO detector is negligible, in agreement with the hitherto non-detection of GW. \\subsection{Limitations and outlook}\\label{lim} The work presented here is a first step towards a detailed study of the dynamics of stellar-mass BHs in star clusters and the consequences for GW-driven BH mergers, and improvements in several directions are possible. First, we do not consider the initial phase of the cluster here when BHs form through stellar evolution and a more consistent approach would be to begin with a star cluster with a full stellar spectrum and produce the BHs from stellar evolution. Note however that in the present study we have inserted numbers of BHs in our clusters of low-mass stars similar to what would have been formed from the stellar evolution of a cluster following a Kroupa IMF (see Sec.~\\ref{runs}). Stellar evolution also produces NSs (about twice in number as the BHs) which also segregate to the central region of the cluster. It is interesting to study the dynamics of the NS cluster and how it is affected by the (more concentrated) BH cluster --- in particular the formation of tight inspiralling NS-NS and NS-BH binaries, which are important for both GW detection and GRBs. Another aspect that we do not consider in our present study is the effect of primordial stellar binaries. Tight stellar binaries aid the formation of compact binaries through double-exchanges \\citep{gr2006}, in addition to the 3-body mechanism (see Sec.~\\ref{intro}) which can increase the number BH-BH (also BH-NS and NS-NS) binaries formed and hence the merger rate. Thus the study of BH-BH binaries in star clusters with primordial binaries is an important next step. Such studies are in progress and will be presented in future papers. Finally, the number of N-body computations in this initial study is not enough to obtain the BH-BH merger rate as a function of cluster mass and BH retention fraction with a reasonable accuracy. There are typically one integration per cluster model. To obtain merger rate dependencies with cluster parameters, \\eg, mass, half mass radius, binary fraction and BH retention, many computations are needed within smaller intervals, which involves much larger time and computing capacity than that utilized for the present project. Such results, combined with improved knowledge of cluster parameter distributions and BH formation through supernova explosions of massive stars, would provide a robust estimate of the BH-BH merger detection rate. Conversely, with such a merger rate function, the detection of BH-BH mergers by GW detectors like AdLIGO in the near future will shed light on the above mentioned long-standing questions." }, "0910/0910.3486_arXiv.txt": { "abstract": "{Earth-sized planets around nearby stars are being detected for the first time by ground-based radial velocity and space-based transit surveys. This milestone is opening the path toward the definition of missions able to directly detect the light from these planets, with the identification of bio-signatures as one of the main objectives. In that respect, both the European Space Agency (ESA) and the National Aeronautics and Space Administration (NASA) have identified nulling interferometry as one of the most promising techniques. The ability to study distant planets will however depend on the amount of exozodiacal dust in the habitable zone of the target stars.} {We assess the impact of exozodiacal clouds on the performance of an infrared nulling interferometer in the Emma X-array configuration. The first part of the study is dedicated to the effect of the disc brightness on the number of targets that can be surveyed and studied by spectroscopy during the mission lifetime. In the second part, we address the impact of asymmetric structures in the discs such as clumps and offset which can potentially mimic the planetary signal.} {We use the \\darwinsim{} software which was designed and validated to study the performance of space-based nulling interferometers. The software has been adapted to handle images of exozodiacal discs and to compute the corresponding demodulated signal.} {For the nominal mission architecture with 2-m aperture telescopes, centrally symmetric exozodiacal dust discs about 100 times denser than the solar zodiacal cloud can be tolerated in order to survey at least 150 targets during the mission lifetime. Considering modeled resonant structures created by an Earth-like planet orbiting at 1\\,AU around a Sun-like star, we show that this tolerable dust density goes down to about 15 times the solar zodiacal density for face-on systems and decreases with the disc inclination.} {Whereas the disc brightness only affects the integration time, the presence of clumps or offset are more problematic and can seriously hamper the planet detection. The upper limits on the tolerable exozodiacal dust density derived in this paper must be considered as rather pessimistic, but still give a realistic estimation of the typical sensitivity that we will need to reach on exozodiacal discs in order to prepare the scientific programme of future Earth-like planet characterisation missions.} ", "introduction": "\\label{sec:intro} The possibility of identifying habitable worlds and even biosignatures from extrasolar planets currently contributes to the growing interest about their nature and properties. Since the first planet discovered around another solar-type star in 1995 \\citep{Mayor:1995}, nearly 370 extrasolar planets have been detected and many more are expected to be unveiled by ongoing or future search programmes. Most extrasolar planets detected so far have been identified from the ground by indirect techniques, which rely on observable effects induced by the planet on its parent star. From the ground, radial velocity measurements are currently limited to the detection of planets about 2 times as massive as the Earth in orbits around Sun-like and low-mass stars \\citep{Mayor:2009b} while the transit method is limited to Neptune-sized planets \\citep{Gillon:2007}. Thanks to the very high precision photometry enabled by the stable space environment, the first space-based dedicated missions (namely CoRoT and Kepler) are now expected to reveal Earth-sized extrasolar planets by transit measurements as well. Launched in 2006, CoRoT (Convection Rotation and planetary Transits) has detected its first extrasolar planets \\citep[e.g.,][]{Barge:2008,Alonso:2008,Leger:2009} and is expected to unveil about 100 transiting planets down to a size of 2\\,R$_{\\oplus}$ around G0V stars and 1.1\\,R$_{\\oplus}$ around M0V stars over its entire lifetime for short orbital periods \\citep{Moutou:2005}. Launched in 2009, Kepler will extend the survey to Earth-sized planets located in the habitable zone of about 10$^5$ main sequence stars \\citep{Borucki:2007}. After 4 years, Kepler should have discovered several hundred of terrestrial planets with periods between one day and 400 days. After this initial reconnaissance by CoRot and Kepler, the Space Interferometry Mission (SIM PlanetQuest) might provide unambiguously the mass of Earth-sized extrasolar planets orbiting in the habitable zone of nearby stars by precise astrometric measurements. With CoRoT and Kepler, we will have a large census of Earth-sized extrasolar planets and their occurrence rate as a function of various stellar properties. However, even though the composition of the upper atmosphere of transiting extrasolar planets can be probed in favorable cases \\citep[e.g.,][]{Richardson:2007}, none of these missions will directly detect the photons emitted by the planets which are required to study the planet atmospheres and eventually reveal the signature of biological activity. Detecting the light from an Earth-like extrasolar planet is very challenging due to the high contrast ($\\sim$10$^7$ in the mid-IR, $\\sim$10$^{10}$ in the visible) and the small angular separation ($\\sim$ 0.5 $\\mu$rad for an Earth-Sun system located at 10\\,pc) between the planet and its host star. A technique that has been proposed to overcome these difficulties is nulling interferometry \\citep{Bracewell:1978}. The basic principle is to combine the beams coming from two telescopes in phase opposition so that a dark fringe appears on the line of sight, which strongly reduces the stellar emission. Considering the two-telescope interferometer initially proposed by Bracewell, the response on the plane of the sky is a series of sinusoidal fringes, with angular spacing of $\\lambda/b$. By adjusting the baseline length ($b$) and orientation, the transmission of the off-axis planetary companion can then be maximised. However, even when the stellar emission is sufficiently reduced, it is generally not possible to detect Earth-like planets with a static array configuration, because their emission is dominated by the thermal contribution of warm dust in our solar system as well as around the target stars (exozodiacal cloud). This is the reason why Bracewell proposed to rotate the interferometer so that the planetary signal is modulated by alternatively crossing high and low transmission regions, while the stellar signal and the background emission remain constant. The planetary signal can then be retrieved by synchronous demodulation. However, a rapid rotation of the array would be difficult to implement and as a result the detection is highly vulnerable to low frequency drifts in the stray light, thermal emission, and detector gain. A number of interferometer configurations with more than two collectors have then been proposed to perform faster modulation and overcome this problem by using phase chopping \\citep{Angel:1997,Mennesson:1997,Absil:2001}. The principle of phase chopping is to synthetize two different transmission maps with the same telescope array, by applying different phase shifts in the beam combination process. By differencing two different transmission maps, it is possible to isolate the planetary signal from the contributions of the star, local zodiacal cloud, exozodiacal cloud, stray light, thermal, or detector gain. Phase chopping can be implemented in various ways (e.g.\\ inherent and internal modulation, \\citealt{Absil:thesis}), and are now an essential part of future space-based life-finding nulling interferometry missions such as ESA's \\darwin{} \\citep{Fridlund:2006} and NASA's Terrestrial Planet Finder \\citep[TPF,][]{Lawson:2008}. The purpose of this paper is to assess the impact of exozodiacal dust discs on the performance of these missions. After describing the nominal performance of \\darwin/TPF, the first part of the study is dedicated to centrally symmetric exozodiacal discs which are suppressed by phase chopping and only contribute through their shot noise. If they are too bright, exozodiacal discs can drive the integration time and we investigate the corresponding impact on the number of planets that can be surveyed during the mission lifetime. In the second part, we address the impact of asymmetric structures in the discs (such as clumps and offset) which are not canceled by phase chopping and can seriously hamper the planet detection process. ", "conclusions": "Infrared nulling interferometry is the core technique of future life-finding space missions such as ESA's \\darwin{} and NASA's Terrestrial Planet Finder (TPF). Observing in the infrared (6-20 $\\mu$m), these missions will be able to characterise the atmosphere of habitable extrasolar planets orbiting around nearby main sequence stars. This ability to study distant planets strongly depends on exozodiacal clouds around the stars, which can hamper the planet detection. Considering the nominal mission architecture with 2-m aperture telescopes, we show that centrally symmetric exozodiacal dust discs about 100 times denser than the solar zodiacal cloud can be tolerated in order to survey at least 150 targets during the mission lifetime. The actual number of planet detections will then depend on the number of terrestrial planets in the habitable zone of target systems. The presence of asymmetric structures in exozodiacal discs (e.g.,\\ clumps or offset) may be more problematic. While the cloud brightness drives the integration time necessary to disentangle the planetary photons from the background noise, the emission from inhomogeneities are not perfectly subtracted by phase chopping so that a part of the disc signal can mimic the planet. To address this issue, we consider modeled resonant structures produced by an Earth-like planet and introduce the corresponding image into \\darwinsim, the mission science simulator. Even for exozodiacal discs a few times brighter than the solar zodiacal cloud, the contribution of these asymmetric structures can be much larger than the planetary signal at the output of the interferometer. Fortunately, the high angular resolution provided by the long imaging baseline of \\darwin/TPF in the X-array configuration is sufficient to spatially distinguish most of the extended exozodi emission from the planetary signal and only the hole in the dust distribution near the planet significantly contributes to the noise level. Considering the full wavelength range of \\darwin/TPF, we show that the tolerable dust density is about 15 times the solar zodiacal density for face-on systems and decreases with the disc inclination. In practice, this constraint might be relaxed since we examined a resonant ring model that does not include dust from highly eccentric or inclined parent bodies, the effects of grain-grain collisions, or perturbations by additional planets, all of which can reduce the contrast of the resonant ring and improve the tolerance to the exozodiacal dust density. These results show that asymmetric structures in exozodiacal discs around nearby main sequence stars are one of the main noise sources for future exo-Earth characterization missions. A first solution to get around this issue is to have a long imaging baseline architecture which resolves out the more spatially extended emission of the exozodiacal cloud from the point-like emission of planets. The stretched X-array configuration is particularly convenient in that respect. The second solution is to observe in advance the nearby main sequence stars and remove from the \\darwin/TPF target list those presenting a too high dust density or disc inclination. The FKSI nulling interferometer would be ideal in that respect with the possibility to detect exozodiacal discs down to the density of the solar zodiacal cloud. Ground-based nulling instruments like LBTI and ALADDIN would also be particularly valuable. \\begin{figure}[!t] \\begin{center} \\includegraphics[width=8.5 cm]{zodi_vs_off.eps} \\caption{Tolerable exozodiacal dust density with respect to the offset between the center of symmetry of the exozodiacal disc and the central star (a G2V star located at 15\\,pc). The disc is assumed to be seen in face-on orientation.}\\label{Fig:zodi_vs_off} \\end{center} \\end{figure} \\appendix" }, "0910/0910.1356_arXiv.txt": { "abstract": "The existence of black holes of masses $\\sim 10^2$--$10^5 {\\rm M_{\\odot}}$ has important implications for the formation and evolution of star clusters and supermassive black holes. One of the strongest candidates to date is the hyperluminous X-ray source HLX1, possibly located in the S0-a galaxy ESO\\,243-49, but the lack of an identifiable optical counterpart had hampered its interpretation. Using the {\\it Magellan} telescope, we have discovered an unresolved optical source with $R = 23.80\\pm0.25$ mag and $V = 24.5\\pm0.3$ mag within HLX1's positional error circle. This implies an average X-ray/optical flux ratio $\\sim 500$. Taking the same distance as ESO\\,243-49, we obtain an intrinsic brightness $M_R = -11.0 \\pm 0.3$ mag, comparable to that of a massive globular cluster. Alternatively, the optical source is consistent with a main-sequence M star in the Galactic halo (for example an M4.4 star at $\\approx 2.5$ kpc). We also examined the properties of ESO\\,243-49 by combining {\\it Swift}/UVOT observations with stellar population modelling. We found that the overall emission is dominated by a $\\sim 5$ Gyr old stellar population, but the UV emission at $\\approx 2000$ \\AA\\ is mostly due to ongoing star-formation at a rate of $\\sim 0.03 {\\rm M_{\\odot}}$ yr$^{-1}$. The UV emission is more intense (at least a $9\\sigma$ enhancement above the mean) North East of the nucleus, in the same quadrant as HLX1. With the combined optical and X-ray measurements, we put constraints on the nature of HLX1. We rule out a foreground star and a background AGN. Two alternative scenarios are still viable. HLX1 could be an accreting intermediate mass black hole in a star cluster, which may itself be the stripped nucleus of a dwarf galaxy that passed through ESO\\,243-49, an event which might have caused the current episode of star formation. Or, it could be a neutron star in the Galactic halo, accreting from an M4--M5 donor star. ", "introduction": "{\\it XMM-Newton} and {\\it Chandra} have discovered several non-nuclear X-ray sources in nearby galaxies, with luminosities up to two orders of magnitude higher than those observed from Galactic X-ray binaries. These are referred to as ultraluminous X-ray sources \\citep[ULXs; e.g.,][]{ggs03,swa04,rob07}. Those findings have challenged our current models of black hole (BH) formation and accretion. Isotropic, Eddington-limited luminosities $\\ga 10^{40}$erg s$^{-1}$ would require BH masses $\\ga 100 {\\rm M_{\\odot}}$, beyond the upper limit for individual stellar collapses \\citep{yun08}. Mildly super-Eddington luminosity (possibly due to large super-Eddington mass accretion) from particularly heavy stellar BHs ($M \\sim 50 {\\rm M_{\\odot}}$), associated with mildly anisotropic emission may explain X-ray luminosities up to $\\sim$ few $\\times 10^{40}$erg s$^{-1}$ without the need for more exotic astrophysical processes \\citep{pou07,rob07,kin09}. Only a few non-nuclear sources have been observed at X-ray luminosities $\\approx 0.7$--$1 \\times 10^{41}$erg s$^{-1}$. For example, in the Cartwheel \\citep{wol07}, in M\\,82 \\citep{fen09}, in NGC\\,2276 \\citep{dav04}, and in NGC\\,5775 \\citep{li08}. It is possible that such rare, extreme ULXs (sometimes known as hyperluminous X-ray sources, HLXs) may be powered by heavier BHs, formed through different channels: for example, in the collapsed core of a super star cluster, or within the nuclear star cluster of an accreted (and now disrupted) dwarf galaxy \\citep{kin05,bek03}. Thus, HLXs may represent evidence of intermediate-mass BHs. However, the debate is far from settled, given the small number of HLXs known, and the possibility of confusion with background AGN. The strongest claim for an X-ray luminous intermediate-mass BH so far has been made for a recently discovered X-ray source \\citep[2XMM J011028.1$-$460421, hereafter HLX1:][]{far09,god09} apparently located in the galaxy ESO\\,243-49, or, at least, projected inside the $\\mu_{B} = 25.0$ mag arcsec$^{-2}$ surface brightness isophote of that galaxy. Here, we report our discovery of the likely optical counterpart to this source, and our analysis of the UV emission in ESO\\,243-49. By determining the optical flux, and the X-ray/optical flux ratio, we test alternative models for the nature of this object. Our results strengthen the interpretation that the X-ray source belongs to ESO\\,243-49. We suggest that it is located inside a massive star cluster. ESO\\,243-49 is an edge-on S0-a galaxy at a luminosity distance of $91 \\pm 6$ Mpc \\citep[$z=0.0224$, distance modulus $34.80 \\pm 0.15$:][]{cal97}. The foreground extinction is very low, $A_V = 0.043$ mag \\citep{sch98}. HLX1 appears projected $\\approx 7\\arcsec$ ($\\approx 3.1$ kpc) to the North-East of ESO\\,243-49's nucleus, and $\\approx 1\\farcs8$ ($\\approx 800$ pc) above the galactic plane. HLX1 has been detected several times with {\\it XMM-Newton}, {\\it Chandra} and {\\it Swift} between 2004 and 2009 \\citep{far09,god09}, with an unabsorbed luminosity in the $0.3$--$10$ keV band varying between $\\la 5 \\times 10^{40}$ erg s$^{-1}$ and $\\approx 1 \\times 10^{42}$ erg s$^{-1}$. We also examined a {\\it ROSAT}/HRI observation of the field from 1996, when HLX1 was not detected to an upper limit of $\\approx 5 \\times 10^{40}$ erg s$^{-1}$. Its combination of extreme luminosity (if it really belongs to ESO\\,243-49), soft spectrum, and spectral changes on short timescales \\citep{god09} makes it a unique object among the ULX/HLX class. Its apparent location in an S0 galaxy is also puzzling, because such galaxies are usually dominated by an old stellar population. For example, the integrated brightnesses of ESO\\,243-49 are (Cousins) $B=14.92 \\pm 0.09$ mag, (Cousins) $R=13.48 \\pm0.09$ mag and (2MASS) $K = 10.70 \\pm 0.05$ mag (from NED\\footnote{http://nedwww.ipac.caltech.edu}), which are indicative of a characteristic stellar age $\\sim$ a few Gyr (Section 4). Such moderately old populations were not previously known to host luminous ULXs or HLXs. For these reasons, it was speculated that the source might be a background AGN or a foreground neutron star, even though its X-ray properties are also very unusual for both classes of objects (Section 5). \\begin{figure} \\psfig{figure=f1_xv.ps,width=84mm} \\caption{{\\it Magellan}/IMACS $R$-band image of ESO\\,243-49, with the X-ray position of HLX1 marked by a circle of $0\\farcs5$ radius (combined astrometric uncertainty of the X-ray and optical images).} \\label{f1} \\end{figure} \\begin{table} \\begin{center} \\begin{tabular}{llrr} \\hline Telescope & Date & Band & Exposure \\\\ \\hline {\\it Magellan} Baade & 2009 Aug 26 & R & 540 s \\\\ & & V & 540 s \\\\ \\hline {\\it Swift}/UVOT & 2008 Oct 24 & $u$ & 380 s\\\\ & & $uvw1$ & 760 s \\\\ & & $uvw2$ & 196 s \\\\ & 2008 Oct 25 & $uvw2$ & 1264 s\\\\ & 2008 Nov 01 & $u$ & 730 s\\\\ & & $uvw1$ & 1690 s \\\\ & & $uvw2$ & 2639 s \\\\ & 2008 Nov 07 & $u$ & 1210 s\\\\ & & $uvw1$ & 2410 s \\\\ & & $uvw2$ & 3278 s \\\\ & 2008 Nov 08 & $uvw2$ & 582 s\\\\ & 2008 Nov 14 & $u$ & 980 s\\\\ & & $uvw1$ & 1960 s \\\\ & & $uvw2$ & 3814 s \\\\ & 2009 Aug 05 & $uvw2$ & 9753 s\\\\ & 2009 Aug 06 & $uvw2$ & 9132 s\\\\ & 2009 Aug 16 & $uvw2$ & 5664 s\\\\ & 2009 Aug 17 & $uvw2$ & 681 s\\\\ & 2009 Aug 18 & $uvw2$ & 6032 s\\\\ & 2009 Aug 19 & $uvw2$ & 4257 s\\\\ & 2009 Aug 20 & $uvw2$ & 2199 s\\\\ & 2009 Nov 02 & $uvw2$ & 8956 s\\\\ & 2009 Nov 14 & $uvw2$ & 3903 s\\\\ & 2009 Nov 20 & $uvw2$ & 517 s\\\\ & 2009 Nov 21 & $uvw2$ & 1453 s\\\\ & 2009 Nov 28 & $uvw2$ & 2981 s\\\\ & 2009 Nov 29 & $uvw2$ & 2028 s\\\\ & 2009 Dec 05 & $uvw2$ & 2993 s\\\\ & 2009 Dec 19 & $uvw2$ & 2518 s\\\\ & 2009 Dec 26 & $uvw2$ & 2774 s\\\\ & 2010 Jan 02 & $uvw2$ & 2590 s\\\\ & 2010 Jan 08 & $uvw2$ & 4348 s\\\\ & 2010 Jan 13 & $uvw2$ & 3577 s\\\\ & 2010 Jan 15 & $uvw2$ & 3068 s\\\\ & 2010 Jan 22 & $uvw2$ & 2945 s\\\\ \\hline \\end{tabular} \\end{center} \\caption{Optical/UV observation log. } \\end{table} \\begin{figure*} \\psfig{figure=new_f2a_xv.ps,width=87mm} \\psfig{figure=new_f2b.ps,width=87mm}\\\\ \\psfig{figure=new_f2c.ps,width=87mm} \\psfig{figure=new_f2d.ps,width=87mm} \\caption{{\\it Top left:} differential $R$-band image of ESO\\,243-49, with a median-filter smoothed image subtracted from the original image (logarithmic greyscale). The source marked with an arrow has a $4\\sigma$ significance and is located near the X-ray position of HLX1. {\\it Top right:} zoomed-in view of the field around HLX1, in the $R$-band residual image (square-root greyscale). The {\\it Chandra}/HRC-I position of HLX1 is marked by a circle of $0\\farcs5$ radius (combined astrometric uncertainty of the X-ray and optical images). Both the top and bottom panel images have been Gaussian-smoothed with a kernel radius of 2 pixels ($0\\farcs22$), for display purposes only. {\\it Bottom left:} Gaussian-smoothed field around HLX1, in the $V$-band residual image (logarithmic greyscale); an optical counterpart is located at the same position with a $3\\sigma$ significance. {\\it Bottom right:} Gaussian-smoothed field around HLX1, from the combined $R$-band plus $V$-band residual images.} \\label{f2} \\end{figure*} ", "conclusions": "If the X-ray source HLX1 is proven to be an accreting BH with mass $\\sim 10^3$--$10^4 {\\rm M_{\\odot}}$, there would be important implications on models of galaxy formation and evolution. Identifying its optical counterpart gives a crucial constraint on its nature. We have found an unresolved optical source within its X-ray error circle, and it is likely to be physically associated to HLX1. Assuming a direct association, we calculate an X-ray/optical flux ratio, using the standard definition \\citep{hor01}: $\\log(f_{\\rm X}/f_{\\rm R}) = \\log f_{\\rm X} + 5.5 + R/2.5$, where $f_{\\rm X}$ is the intrinsic flux in the $0.3$--$10$ keV band \\citep[taken from][and our spectral analysis of the {\\it Swift}/XRT data]{god09}. We obtain $f_{\\rm X}/f_{\\rm R} \\approx 800$--$1000$ for the X-ray high state of 2009 August, and $f_{\\rm X}/f_{\\rm R} \\approx 500$ for the ``average'' X-ray state where the source was more often observed over 2008--2009; such ratio is only slightly dependent on the choice of X-ray spectral model. The observed X-ray/optical flux ratio is much higher than expected for AGN, which have typical $f_{\\rm X}/f_{\\rm R} \\la 10$ \\citep{lai09,bau04}. A number of distant, faint AGN are undetected in the optical band because of extinction, which is not an issue for this object \\citep{far09,god09}. In particular, AGN with a $0.5$--$2$ keV flux of $\\approx 5 \\times 10^{-13}$ erg s$^{-1}$ cm$^{-2}$ are rare \\citep[$N \\approx 0.5$ deg$^{-2}$:][]{has98} and have always been easily identified in other bands. Moreover, the red colour of the optical counterpart ($V-R = 0.7 \\pm 0.4$) suggests that the optical emission is not dominated by the Rayleigh-Jeans tail of an accretion disk spectrum. Neutron stars are a class of objects that can reach X-ray/optical flux ratios $\\ga 1000$, with thermal X-ray emission from their surface. We note (Soria et al., in prep.) that the X-ray spectra of HLX1 can also be fitted with a neutron star atmosphere model (e.g., {\\tt nsa} in {\\footnotesize XSPEC}) plus power-law. Such models suggest a characteristic distance $\\approx 1.5$--$2.5$ kpc, placing the source in the Galactic halo. The corresponding $0.3$--$10$ keV luminosity would be $\\approx$ a few $\\times 10^{32}$ erg s$^{-1}$. Intriguingly, this is a range of luminosities where quiescent low-mass X-ray binaries also show a thermal plus power-law spectrum, with the relative contribution of the power-law decreasing as the source gets brighter \\citep{jon04}. In this scenario, the apparent optical brightness $R \\approx 23.8$ mag implies $M_R \\approx 11.8$--$12.8$ mag, consistent with a main-sequence M4.4--M5.2 donor star, with an initial mass $\\approx 0.13$--$0.17 {\\rm M_{\\odot}}$ \\citep{kni06,gir00}. A late-type M star is also consistent with the observed red colour. Thus, we cannot rule out the possibility of an accreting neutron star in the Galactic halo, from the optical data. In this case, the excess UV emission North East of the nucleus in ESO\\,243-49 is purely coincidental. The optical counterpart of HLX1 does not stick out like a unique object in the field. In the $R$-band residual image, we identified a few other sources consistent with globular clusters around ESO\\,243-49, with comparable brightnesses (Figure 2, top panel). If the HLX1 counterpart is also a globular cluster in that galaxy, its optical luminosity would place it between the Milky Way globular cluster $\\omega$ Cen \\citep[$M_V = -10.3$ mag, $M_R = -10.8$ mag, $M_{\\rm tot} \\approx 2.8 \\times 10^6 {\\rm M_{\\odot}}$:][]{har96,van09} and the Andromeda cluster G1 \\citep[$M_V = -11.2$ mag, $M_R = -11.8$ mag, $M_{\\rm tot} \\approx 5 \\times 10^6 {\\rm M_{\\odot}}$:][]{gra09}, which is a strong candidate for the presence of an $\\approx 2 \\times 10^4 {\\rm M_{\\odot}}$ BH \\citep[][but see \\citealt{bau03}]{geb05}. There are a number of scenarios for the formation of an intermediate-mass BH inside a massive star cluster. In a young star cluster, runaway core collapse and coalescence of the most massive stars can occur over a timescale $\\la 3$ Myrs and can result in the formation of a supermassive star, which can quickly collapse into a BH \\citep{por02,fre06}. In an old globular cluster, an intermediate-mass BH can be formed from the merger of stellar-mass BHs and neutron stars over a timescale $\\sim 10^9$ yr \\citep{ole06}. Subsequent capture and disruption of a cluster star may then provide the accretion rate required to explain the X-ray luminosity. In this scenario, HLX1 and its host star cluster are unrelated to the ongoing star formation in the bulge of ESO\\,243-49. The most intriguing scenario is that some massive star clusters may have been the nuclear clusters of satellite galaxies accreted and tidally disrupted by a more massive galaxy. Dwarf galaxies are the most common type of galaxies in clusters \\citep[e.g.,][]{bin85} and many of them are nucleated \\citep[e.g.,][]{gra03,cot06}. In many cases, a nuclear cluster may coexist with a nuclear BH \\citep{gra09,set08}. This may end up in the halo of a bigger galaxy after a merger. $\\omega$ Cen itself may have originated from the nuclear star cluster of an accreted dwarf \\citep{bek03}. Similar suggestions have been made for a group of clusters in NGC\\,5128 \\citep{pen02,cha09}. The recent or ongoing star formation in ESO\\,243-49 may have been triggered by the passage and tidal disruption of the satellite galaxy, perhaps along the South-West to North-East direction, since {\\it uvw2} emission is stronger on that side (Section 4). During its passage through ESO\\,243-49, the compact nucleus of the satellite dwarf may also have collected gas from the main galaxy \\citep{pfl09}, and this may perhaps be fuelling a nuclear BH. In this scenario, HLX1 may be the intermediate-mass BH located in the nuclear cluster of that accreted satellite. In summary, we have identified a point-like optical counterpart for HLX1 in ESO\\,243-49 in the {\\it Magellan} images. The optical brightness and colour are consistent with a massive star cluster in ESO\\,243-49, or with a main-sequence M4--M5 star in the Milky Way halo, at $\\approx 1.5$--$2.5$ kpc. The galaxy is dominated by a $\\sim 5$ Gyr old population, but shows excess emission in the {\\it Swift}/UVOT $uvw2$ band, consistent with a recent episode of star formation. The far-UV emission has an asymmetric shape and is stronger to the North East of the nucleus, roughly in the direction of HLX1." }, "0910/0910.4745_arXiv.txt": { "abstract": "Particle cascades initiated by ultra-high energy (UHE) neutrinos in the lunar regolith will emit an electromagnetic pulse with a time duration of the order of nano seconds through a process known as the Askaryan effect. It has been shown that in an observing window around 150~MHz there is a maximum chance for detecting this radiation with radio telescopes commonly used in astronomy. In 50 hours of observation time with the Westerbork Synthesis Radio Telescope array we have set a new limit on the flux of neutrinos, summed over all flavors, with energies in excess of $4\\times10^{22}$~eV. ", "introduction": "At high energies the spectrum of cosmic rays follows a power law distribution extending up to extremely large energies. At the Pierre Auger Observatory cosmic rays have been observed~\\cite{A08} with energies in excess of $\\sim 10^{20}$~eV. Above the Greisen-Zatsepin-Kuzmin (GZK) energy of $6\\times 10^{19}$~eV, cosmic rays can interact with the photons of the cosmic microwave background to produce pions~\\cite{G66,ZK66} which carry a sizable fraction of the original energy of the cosmic ray. Charged pions decay and produce neutrinos and one thus may expect the presence of neutrinos with energies in excess of the GZK energy. Recently at the Pierre Auger Observatory a steepening of the slope in the cosmic ray spectrum has been observed at the GZK energy~\\cite{A08} which can be regarded as a consequence of the GZK effect. Since neutrinos are chargeless they will propagate in a straight line with negligible energy loss from the location where they have been created to the observer, thus carrying direct information on their source. These sources could be the aforementioned processes related to the GZK effect or, more exotically, decaying supermassive dark-matter particles or topological defects. This last class of models is referred to as top-down (see \\citet{s04} for a review). Because of their small interaction cross section, the detection of cosmic neutrinos calls for extremely large detectors. At GeV energies their cross section is so minute that at a flux given by the Waxman-Bahcall estimate~\\cite{Waxman:1998yy} of a few tens of neutrinos per km$^2$ per year one needs to employ km$^3$-scale detectors~\\cite{icecube,KM3NET}. At higher neutrino energies the reaction cross section increases. However, their flux is expected to fall even faster and one needs even larger detection volumes. These can be obtained by observing large detector masses from a distance. The ANITA balloon mission~\\cite{anita08} monitors an area of a million km$^2$ of South Pole ice and the FORTE satellite~\\cite{forte} can pick up radio signals coming from the Greenland ice mass. Alternatively the Pierre Auger Observatory can distinguish cosmic ray induced air showers from neutrino induced cascades at very high zenith angles~\\cite{Abraham:2007rj}. The Moon offers an even larger natural detector volume. In the interaction of an UHE neutrino about 20\\% of the energy of the neutrino is converted into a cascade of high-energy particles, called the hadronic shower. Due to the electromagnetic component of this shower the electrons in the material are swept out from the atom to become part of the shower. The shower thus has an excess of negative charge moving with relativistic velocities through a material with an index of refraction considerably different from unity resulting in the emission of Cherenkov radiation. Since the lateral side of the shower has a dimension of the order of 10~cm, the radio emission is coherent for wavelength of this magnitude and larger or for frequencies up to $\\sim 3$ GHz. The emission of coherent Cherenkov radiation in such a process is known as the Askaryan effect~\\cite{a62}. This emission mechanism has been experimentally verified at accelerators~\\cite{s01} and extensive calculations have been performed to quantify the effect~\\cite{zhs92,az97}. For showers in the lunar regolith, the top layer of the Moon consisting of dust and small rocks, its properties are important. Much is known from samples brought from the Moon~\\cite{os75}. The average index of refraction is $n=1.8$ and the attenuation length is $\\lambda_{r}=(9/\\nu[\\mathrm{GHz}])$~m for radio waves. The thickness of the regolith is known to vary over the lunar surface. At some depth there is a (probably smooth) transition to solid rock, for which the density is about twice that of the regolith. At the highest energies the emitted pulse from the Moon can be observed at Earth with radio telescopes~\\cite{dz89}. The first experiments in this direction were carried out with the Parkes telescope~\\cite{parkes}, later followed by others~\\cite{glue,KALYAZIN}. A recent project is LUNASKA~\\cite{James:2009rc} that is currently performing lunar Cherenkov measurements with the Australia Telescope Compact Array with a 600 MHz bandwidth at 1.2-1.8~GHz. These observations are all performed at relatively high frequencies (2~GHz) where the emission is strongest. At lower frequencies~\\cite{Fal03} the angular spread of the emission around the Cherenkov angle increases due to finite source effects~\\cite{scholten}. When the wavelength is similar to the longitudinal extent of the shower in the lunar rock, a few meters, the angular spread is close to isotropic and the probability of detecting the radio pulse is largest~\\cite{scholten}. In our observations we exploit this optimal frequency range around 150 MHz using the Westerbork Synthesis Radio Telescope (WSRT). ", "conclusions": "" }, "0910/0910.3023_arXiv.txt": { "abstract": "We present initial results from a new 440-ks Chandra HETG GTO observation of the canonical Seyfert 2 galaxy NGC 1068. The proximity of NGC 1068, together with Chandra's superb spatial and spectral resolution, allow an unprecedented view of its nucleus and circumnuclear NLR. We perform the first spatially resolved high-resolution X-ray spectroscopy of the `ionization cone' in any AGN, and use the sensitive line diagnostics offered by the HETG to measure the ionization state, density, and temperature at discrete points along the ionized NLR. We argue that the NLR takes the form of outflowing photoionized gas, rather than gas that has been collisionally ionized by the small-scale radio jet in NGC 1068. We investigate evidence for any velocity gradients in the outflow, and describe our next steps in modeling the spatially resolved spectra as a function of distance from the nucleus. ", "introduction": "Outflows and feedback from AGN are widely invoked as the key mediator between the co-evolution of black holes and their host galaxies over cosmic time. It is now well understood from large optical surveys that galaxies evolve through mergers from blue, star-forming spirals (the `blue cloud') whose black holes accrete at or near their Eddington limits, through a transition region (the `green valley'), to so-called `red and dead' ellipticals (the `red sequence'), which instead are described by markedly less black-hole growth. The importance of outflows in this evolution has now taken center stage; yet before we can successfully incorporate AGN feedback into numerical simulations of galaxy growth, a number of key questions need to be answered from observations: \\begin{itemize} \\item Can AGN actually deliver enough power to their environments to alter the evolution of their host galaxies in a meaningful way? \\item On what spatial scales does this occur? \\item Where are the outflows in the first place? \\end{itemize} Several steps have recently been taken to address these issues. It is now apparent from multiwavelength surveys that AGN selected in different wavelengths tend to populate different regions of the galaxy color-magnitude diagram. For example, \\cite{hic09} showed that IR-selected AGN are characterized by high Eddington ratios and tend to inhabit the `blue cloud'; X-ray AGN have intermediate-to-high ratios and lie in the `green valley'; and radio AGN have low Eddington fractions and populate the `red sequence'. The relative paucity of galaxies in the `green valley', together with the prevalence of X-ray AGN in this region suggest that the kpc-scale, ionized Narrow-Line Regions (NLR) in such AGN are ideal places to search for outflows, together with their kinematic and ionizing effect on their galaxy-scale environments. ", "conclusions": "" }, "0910/0910.0216_arXiv.txt": { "abstract": "We investigate the solar flare of 20 October 2002. The flare was accompanied by quasi-periodic pulsations (QPP) of both thermal and nonthermal hard X-ray emissions (HXR) observed by RHESSI in the 3-50 keV energy range. Analysis of the HXR time profiles in different energy channels made with the Lomb periodogram indicates two statistically significant time periods of about 16 and 36 seconds. The 36-second QPP were observed only in the nonthermal HXR emission in the impulsive phase of the flare. The 16-second QPP were more pronounced in the thermal HXR emission and were observed both in the impulsive and in the decay phases of the flare. Imaging analysis of the flare region, the determined time periods of the QPP and the estimated physical parameters of magnetic loops in the flare region allow us to interpret the observations as follows. \\textbf{1)} In the impulsive phase energy was released and electrons were accelerated by successive acts with the average time period of about 36 seconds in different parts of two spatially separated, but interacting loop systems of the flare region. \\textbf{2)} The 36-second periodicity of energy release could be caused by the action of fast MHD oscillations in the loops connecting these flaring sites. \\textbf{3)} During the first explosive acts of energy release the MHD oscillations (most probably the sausage mode) with time period of 16 seconds were excited in one system of the flare loops. \\textbf{4)} These oscillations were maintained by the subsequent explosive acts of energy release in the impulsive phase and were completely damped in the decay phase of the flare. ", "introduction": "\\label{S-Introduction} \\indent Electrons, accelerated during solar flares, generate nonthermal HXR and microwave emission in the solar atmosphere as a result of their Coulomb collisions with ambient plasma (bremsstrahlung) and interaction with magnetic field (synchrotron radiation), respectively. Often the light curves of nonthermal HXR and microwave emissions in large solar flares are composed of multiple pulses with different durations from less than one second up to several minutes (\\eg,~ \\opencite{dennis88};~\\opencite{aschwanden09}, and references therein). This indicates many episodes of acceleration of electrons, possibly in different locations within flare regions, since the majority of these flares are accompanied by formation of arcades of many flaring loops and by the HXR emission generated in different flare loops (\\eg,~ \\opencite{vorpahl76};~\\opencite{dejager79};~\\opencite{grigis05}). It has been suggested that flares consist of a number of ``Elementary Flare Bursts'' (EFB), which can be a result of collisions between different current-carrying loops \\cite{dejager79,sakai96}. \\indent More rarely, the light curves of nonthermal HXR and microwave emission show apparent quasi-periodic pulsations (QPP; see, \\eg,~ \\opencite{dennis88}; ~\\opencite{nakariakov05};~ \\opencite{aschwanden09} for a review and references therein). This may indicate the presence of some quasi-periodic processes (not fully understood at the moment) of acceleration of electrons. Using RHESSI imaging observations in the HXR range, it was shown that this quasi-periodic acceleration of electrons, at least in some flares, occurs in different flare loops, rather than in a single one \\cite{zimovets09}. In this case the observed quasi-periodicity could be a coincidence owing to similar physical conditions in successively bursting flare loops. This returns us to the idea of multiple EFB due to coalescence of current-carrying loops or due to the presence of multiple spatially separated ``null-points'' in flare regions, in which the stored magnetic energy can be released. But in principle, it is possible to implement a model proposed by \\inlinecite{nakariakov06} to explain the QPP in the latter case by the successive quasi-periodic reconnection in different null-points due to fast magnetoacoustic oscillations in nearby non-flaring loops. \\indent On the other hand, it was theoretically suggested that QPP could be produced inside a single vibrating flare loop by the quasi-periodic modulation of the spectrum of nonthermal electrons through the betatron acceleration \\cite{brown75} or by the quasi-periodic modulation of efficiency of the trapped nonthermal electrons to precipitate into the lower heights with larger plasma density and stronger magnetic field (in particular, the sausage mode is a good candidate for this) \\cite{zaitsev82}. Indeed, imaging observations of the QPP in the microwave range made using the Nobeyama Radioheliograph with sufficient spatial resolution (but with no simultaneous QPP observed in the HXR range), which could be interpreted in terms of the global sausage mode of the flaring loop oscillations, were reported \\cite{nakariakov03}. Imaging observations of QPP, but without sufficient spatial resolution together with an analysis of the time periods of QPP, allowed QPP to be interpreted in terms of the flare loop oscillations (\\eg,~ \\opencite{asai01};~ \\opencite{inglis09}). Different modes of oscillations in coronal loops actually have been found with the TRACE and SOHO spacecraft in the range of the extreme ultraviolet radiation (see, \\eg,~ \\opencite{aschwanden09};~ \\opencite{nakariakov05} for a review and references therein), making the oscillations of magnetic loops a natural way to explain the QPP of electromagnetic emission in flares. However, despite the large number of papers published on flare QPP, their nature is still not fully understood. Further imaging analysis of flares with QPP is required. \\indent In this work we investigate the solar flare of 20 October 2002. The light curves of its thermal and nonthermal HXR emission detected by RHESSI clearly indicate the presence of the QPP with two significant time periods of about 16 and 36 seconds. We will present time-spectral, energy-spectral and imaging analysis of the flare HXR emission and propose possible flare scenario to interpret the observed QPP. ", "conclusions": "% \\label{S-discussion} Based on the analyzed observational data, we can propose a rough scenario of the investigated flare [see Figure~\\ref{fig3}(p)]. \\indent \\textbf{1.} \\textbf{In the impulsive phase of the flare energy was released and electrons were accelerated by successive quasi-periodic acts with an average time period of about 36 seconds in different parts of two spatially separated, but interacting loop-systems of the flare region.} Possibly, these loop-systems interact through the connecting magnetic loop near both ends of which they are located. We have established two observational evidences that these flare systems are linked: 1) the fluxes of thermal HXR emission from these systems are strongly correlated; 2) the formation of the loop-type structure in the space between these flaring systems is observed in thermal HXR emission. The question appears: is it possible to interpret the observed 36-second quasi-periodicity ($P_{36}$) of energy release in two different flaring systems by the MHD oscillations in the connecting loop? Indeed, it was concluded from the imaging observations of another flare event that oscillations of even nonflaring transequatorial loop (the kink modes) can be responsible for the QPP of HXR emission observed simultaneously from two different active regions located at both ends of this loop \\cite{foullon05}. Unfortunately, it is not possible to estimate length of the connecting loop ($L_{c}$) accurately in our case, due to insufficient spatial resolution of observations, nevertheless we can say with certainty that $20\\leq{L_{c}}\\leq40$ Mm. Thus, we can estimate the required phase speed of a fundamental standing mode as $V_{ph}=2L_{c}/P_{36}$ or $1100\\leq{V_{ph}}\\leq{2200}$ km/s. These phase speeds are consistent with the speeds of fast magnetoacoustic kink modes that have been directly observed in coronal loops in the EUV emission (\\opencite{nakariakov05}; \\opencite{aschwanden09}; and references therein). Implementing the imaging spectroscopy technique to the RHESSI HXR data we could estimate the emission measure of different parts of the flare region. Hence, using the observed volumes of these parts, we could roughly estimate averaged electron plasma density ($n_{e}$) in them. Our estimations give ${10^{10}}\\leq{n_{e}}\\leq{10^{11}}$ cm$^{-3}$ for each flaring loop-system and also for the connecting loop-type structure. Thus, the Alfv\\'{e}n speed in the connecting loop-type structure is $7B\\leq{V_{A}}\\leq{22B}$ ~km/s, where $B$ ~is magnetic field in gauss inside the loop-type structure. Consequently, if $B\\approx{100}$ ~G, which is a reasonable value, we have $700\\leq{V_{A}}\\leq{2200}$ ~km/s. These values of the Alfv\\'{e}n speed are very close to the estimated $V_{ph}$! Thus, we conclude that the observed 36-second periodicity can actually be caused by the fast MHD oscillations in the loop-type structure, which connects two separate flare systems. The physical mechanism, proposed by \\inlinecite{nakariakov06}, is a good candidate to explain quasi-periodic spatial fragmentation of energy release in the impulsive phase of this flare (see Section~\\ref{S-Introduction}). Multiple null-points could actually exist in this flare region, because of its complex topology: several inclusions of magnetic field of opposite polarity are seen on the magnetograms. \\indent \\textbf{2.}\t\\textbf{During the first 36-second explosive episodes of energy release, the MHD oscillations, probably global sausage modes} (see below)\\textbf{, with time period of about 16 seconds were excited in the loops of the NW flare system.} \\indent \\textbf{3.}\t\\textbf{These oscillations were maintained by the subsequent explosive acts of energy release in the impulsive phase and were damped in the decay phase of the flare.} We conclude that the 16-second oscillations were excited only in the NW system, but not in the SE flare system and not in the connecting loop-type structure at least in the decay phase, because the NW system was more stationary and retained its loop-like configuration in the decay phase, when the QPP of thermal HXR emission were still observed, but the SE flare system and the connecting loop-type structure were no longer visible in thermal HXR emission on the RHESSI images. Estimated length and electron plasma density of the loops in the NW flare system are $25\\leq{L_{NW}}\\leq{35}$ Mm and ${10^{10}}\\leq{n_{e}}\\leq{10^{11}}$~cm$^{-3}$, respectively. These physical parameters highly coincide with the same parameters, found in \\inlinecite{nakariakov03}, where the 16-second quasi-periodic pulsations of microwave emission, observed with the Nobeyama Radioheliograph, were successfully interpreted in terms of the global sausage mode of the oscillating flare loop! Thus, by analogy, we can also interpret the observed 16-second pulsations of HXR emission as the global sausage mode, excited in the loops of the NW flare system. Moreover, the sausage mode must be accompanied by perturbations of plasma density and emission measure. Hence, this mode of oscillations would quasi-periodically modulate flux of thermal HXR emission. This is what we observe both in the impulsive and in the decay phase of the flare. The decay phase was not accompanied by significant energy release, therefore we clearly observe the damping of oscillations. Contrary to that, in the impulsive phase there were multiple explosive acts of energy release, which might intermittently re-excite oscillations. \\indent It is interesting to note that the 16-second periodicity of nonthermal emission has already been observed in other solar flares \\cite{parks69,nakariakov03,inglis08}. Perhaps, this is a typical period of the global sausage oscillations in flare regions \\cite{inglis08}. The 36-second periodicity of HXR emission was also found in many solar flares \\cite{lipa78}. \\indent Despite the above arguments we are aware that the proposed flare scenario, based on the MHD oscillations of magnetic loops in flare region, may not be true. Oscillations of the flaring loops itself were not detected directly using imaging observations, though this may be due to the observational limitations. \\begin{acks} We are grateful to the spacecraft teams and consortia (RHESSI, \\\\ SOHO, and GOES) and ground-based observatories (the Sagamore Hill Radio Observatory and the San Vito Solar Observatory), whose data were used in this work. We thank Jenny Harris for help in English editing. IVZ is grateful to the organizers of the Baikal Young Scientists' International School for Fundamental Physics (2009), where this work was presented. This work was partially supported by the Russian Foundation for Basic Research, grants No. 07-02-00319, 09-02-16032. \\indent This article is dedicated to IVZ's grandfather F.I.~Krylov, veteran of the WWII, who died during its writing. \\end{acks} \\clearpage{}" }, "0910/0910.2025_arXiv.txt": { "abstract": "{} {Giant radio halos are mega-parsec scale synchrotron sources detected in a fraction of massive and merging galaxy clusters. Radio halos provide one of the most important pieces of evidence for non-thermal components in large scale structure. Statistics of their properties can be used to discriminate among various models for their origin. Therefore, theoretical predictions of the occurrence of radio halos are important as several new radio telescopes are about to begin to survey the sky at low frequencies with unprecedented sensitivity.} {In this paper we carry out Monte Carlo simulations to model the formation and evolution of radio halos in a cosmological framework. We extend previous works on the statistical properties of radio halos in the context of the turbulent re-acceleration model.} {First we compute the fraction of galaxy clusters that show radio halos and derive the luminosity function of radio halos. Then, we derive differential and integrated number count distributions of radio halos at low radio frequencies with the main goal to explore the potential of the upcoming LOFAR surveys. By restricting to the case of clusters at redshifts $<$ 0.6, we find that the planned LOFAR all sky survey at 120 MHz is expected to detect about 350 giant radio halos. About half of these halos have spectral indices larger than 1.9 and substantially brighten at lower frequencies. If detected they will allow for a confirmation that turbulence accelerates the emitting particles. We expect that also commissioning surveys, such as MS$^3$, have the potential to detect about 60 radio halos in clusters of the ROSAT Brightest Cluster Sample and its extension (eBCS). These surveys will allow us to constrain how the rate of formation of radio halos in these clusters depends on cluster mass.} {} ", "introduction": "Radio halos are diffuse Mpc--scale radio sources observed at the center of $\\sim 30\\%$ of massive galaxy clusters (\\egg Feretti 2005; Ferrari et al. 2008, for reviews). These sources emit synchrotron radiation due to GeV electrons diffusing through $\\mu$G magnetic fields and provide the most important evidence of non-thermal components in the intra-cluster-medium (ICM). Clusters hosting radio halos always show very recent or ongoing merger events (\\eg Buote 2001; Schuecker et al 2001; Govoni et al. 2004; Venturi et al. 2008). This suggests a connection between the gravitational process of cluster formation and the origin of these halos. Cluster mergers are expected to be the most important sources of non-thermal components in galaxy clusters. A fraction of the energy dissipated during these mergers could be channelled into amplification of the magnetic fields (\\eg Dolag et al. 2002; Br\\\"uggen et al. 2005; Subramanian et al. 2006; Ryu et al. 2008) and into the acceleration of high energy particles via shocks and turbulence (\\eg En\\ss lin et al. 1998; Sarazin 1999; Blasi 2001; Brunetti et al. 2001, 2004; Petrosian 2001; Miniati et al. 2001; Fujita et al. 2003; Ryu et al. 2003; Hoeft \\& Br\\\"uggen 2007; Brunetti \\& Lazarian 2007; Pfrommer et al. 2008, Brunetti et al. 2009). A promising scenario proposed to explain the origin of the synchrotron emitting electrons in radio halos assumes that electrons are re-accelerated due to the interaction with MHD turbulence injected in the ICM in connection with cluster mergers ({\\it turbulent re-acceleration} model, \\eg Brunetti et al. 2001; Petrosian 2001). An alternative possibility is that the emitting electrons are continuously injected by {\\it pp} collisions in the ICM ({\\it secondary} models; \\eg Dennison 1980; Blasi \\& Colafrancesco 1999). In the picture of the {\\it turbulent re-acceleration} scenario, the formation and evolution of radio halos are tightly connected with the dynamics and evolution of the hosting clusters. Indeed, the occurrence of radio halos at any redshift depends on the rate of cluster-cluster mergers and on the fraction of the merger energy channelled into MHD turbulence and re-acceleration of high energy particles. In the last few years this has been modeled through Monte Carlo procedures (Cassano \\& Brunetti 2005; Cassano et al. 2006a) that provide predictions that can be verified using future instruments. In this scenario radio halos have a relatively short lifetime ($\\approx$ 1 Gyr), and the fraction of galaxy clusters where radio halos are generated is expected to increase with cluster mass (or X-ray luminosity), since the energy of the turbulence generated during cluster mergers is expected to scale with the cluster thermal energy (which roughly scales as $\\sim M^{5/3}$; \\eg Cassano \\& Brunetti 2005). It has been shown that the predicted occurrence of radio halos as a function of the cluster mass (or X-ray luminosity) is in line with results obtained from a large observational project, the ``GMRT radio halo survey'' (Venturi et al.~2007, 2008), and with its combination with studies of nearby halos based on the NVSS survey (\\eg Cassano et al.~2008). \\begin{figure} \\centerline{ \\includegraphics[width=7.3cm,height=6.5cm]{USSRH_paper2.ps} } \\caption{A schematic representation of the theoretical synchrotron spectra of radio halos with different values of $\\nu_s$ (and $\\nu_b$). The colored regions indicate the frequency range of VLA and LOFAR observations.} \\label{fig:ussrh} \\end{figure} \\noindent The steep spectrum of radio halos makes these sources ideal targets for observations at low radio frequencies suggesting that present radio telescopes can only detect the tip of the iceberg of their population (En\\ss lin \\& R\\\"ottgering 2002; Cassano et al.~2006; Hoeft et al. 2008). The recent discovery of the giant and ultra-steep spectrum radio halo in Abell 521 at low radio frequencies (Brunetti et al. 2008) allows a first confirmation of this conjecture and provides a glimpse of what future low frequency radio telescopes, such as the Low Frequency Array (LOFAR)\\footnote{http://www.lofar.org} and the Long Wavelength Array (LWA, \\eg Ellingson et al. 2009), might find in the upcoming years. \\noindent LOFAR promises an impressive gain of two orders of magnitude in sensitivity and angular resolution over present instruments in the frequency range 15--240 MHz, and as such will open up a new observational window to the Universe. In particular, LOFAR is expected to contribute significantly to the understanding of the origin and evolution of the relativistic matter and magnetic fields in galaxy clusters. The main focus of the present paper is to provide a theoretical framework for the interpretation of future LOFAR data by quantifying expectations for the properties and occurrence of giant radio halos in the context of the {\\it turbulent re-acceleration} scenario. In particular, in Sect. 2 we summarize the main ingredients used in the model calculations and provide an extension of the results of previous papers on the occurrence of radio halos in clusters (Sect.~2.1) and on the expected radio halo luminosity functions (Sect. 2.2). In Sect.~3 we derive the expected number counts of radio halos at 120 MHz and explore the potential of LOFAR surveys. Our conclusions are given in Sect.4. \\noindent A $\\Lambda$CDM ($H_{o}=70$ $\\mathrm{ Km\\, s^{-1} Mpc^{-1}}$, $\\Omega_{m}=0.3$, $\\Omega_{\\Lambda}=0.7$) cosmology is adopted throughout the paper. ", "conclusions": "In the present paper we perform Monte Carlo simulations to model the formation and evolution of giant radio halos in the framework of the merger-induced particle acceleration scenario (see Sec. 2). Following Cassano et al. (2006a) we use homogeneous models that assume {\\it a)} an average value of the magnetic field strength in the radio halo volume that scales with cluster mass as $B = B_{} M_v^b$, and {\\it b)} that a fraction, $\\eta_t$, of the $P dV$ work, done by subclusters crossing the main clusters during mergers goes into {\\it magneto-acoustic} turbulence. Although simple, these models reproduce the presently observed fraction of galaxy clusters with radio halos and the scalings between the monochromatic radio power of halos at 1.4 GHz and the mass and X-ray luminosity of the host clusters (\\eg Cassano et al.~2006, 2008; Venturi et al. 2008), provided that the model parameters $(B_{}, b , \\eta_t)$ lie within a fairly constrained range of values (Fig.~7 in Cassano et al. 2006a); in the present paper we adopt a reference set of parameters: $=1.9\\, \\mu$G, $b=1.5$, $\\eta_t=0.2$, that falls in that range. In Fig.~\\ref{fig:NC_NVSS_GMRT} we show that the expected number counts of giant radio halos at $\\nu_o=1.4$ GHz obtained with this set of parameters are nicely in agreement with both the data at low redshift (NVSS-XBACS selected radio halos, Giovannini et al.~1999) and at intermediate redshift (clusters in the ``GMRT radio halo survey'', Venturi et al.~2007, 2008). The most important expectation of the {\\it turbulent re-acceleration} scenario is that the synchrotron spectrum of radio halos should become gradually steeper above a frequency, $\\nu_s$ that is determined by the energetics of the merger events that generate the halos and by the electron radiative losses (\\eg Fujita et al.~2003; Cassano \\& Brunetti 2005). Consequently, the population of radio halos is expected to be constituted by a mixture of halos with different spectra, with steep spectrum halos being more common in the Universe than those with flatter spectra (\\eg Cassano et al.~2006). The discovery of these very steep spectrum halos will allow to test the above theoretical conjectures. In Sect.~2 we have derived the expected radio halo luminosity functions (RHLF) at frequency $\\nu_o$, that account for the contribution from the different populations of radio halos with $\\nu_s \\geq \\nu_o$. The RHLF are obtained combining the theoretical mass function of radio halos (with different $\\nu_s \\geq \\nu_o$) with the radio power--cluster mass correlation (Eq.\\ref{RHLF}). The expected monochromatic radio power at $\\nu_o$ of halos hosted by clusters with mass $M_v$ is extrapolated from the observed $P(1.4)$--$M_v$ correlation by assuming simple scaling relations, appropriate for homogenous models, that account for the dependence of the emitted synchrotron power on $\\nu_s$ (Eqs.\\ref{scale2},\\ref{scale3}). \\noindent As a relevant case we calculate the expected RHLF at $\\nu_o$= 120 MHz (Fig.~\\ref{Fig.RHLF}). The shape of the RHLF can be approximated by a power law over more than two orders of magnitude in radio power. Homogeneous models imply the following scalings between $\\nu_s$, cluster mass and the radio luminosity at $\\nu_o$, $P_{\\nu_s} (\\nu_o)$ : \\begin{equation} \\nu_s \\propto M^{4/3 +b} {{(1 + \\Delta M / M)^3}\\over {( ^2 + B_{cmb}^2 )^2}} \\label{nus_concl} \\end{equation} \\noindent and from Eq.~\\ref{scale3} and the $P_{1.4}-M_v$ correlation: \\begin{equation} P_{\\nu_s} (\\nu_o) \\propto M_v^3 \\nu_s^{\\alpha} \\label{p_concl} \\end{equation} \\noindent i.e., radio halos with larger $\\nu_s$ are typically generated in massive clusters that undergo major mergers and contribute to the RHLF at larger powers. On the other hand halos with smaller $\\nu_s$ are typically generated in less massive systems and contribute to the RHLF at fainter powers. Radio halos with $\\nu_s \\geq$ 120 MHz however become increasingly rare in clusters with mass $\\leq 5 \\times 10^{14}$M$_{\\odot}$ and this explains the drop of the RHLF at lower radio powers in Fig.~\\ref{Fig.RHLF}. At the same time, halos with monochromatic radio emission at 120 MHz $> 10^{26}$W Hz$^{-1}$ would be generated in connection with very energetics merging events in very massive clusters, that are extremely rare, and this explains the RHLF cut-off at higher synchrotron powers in Fig.~\\ref{Fig.RHLF}. In Sect.~3 we discuss the expected number counts of radio halos at 120 MHz, that best allow us to explore the potential of incoming LOFAR surveys in constraining present models. \\noindent The crucial step in this analysis is the estimate of the minimum diffuse flux from giant radio halos that is detectable by these surveys. Because the LOFAR capabilities will become more clear during the incoming commissioning phase we exploit two complementary approaches: {\\it i)} we required that at least half of the radio halo emission is above a fixed brightness--threshold, $\\xi\\,F$ ($F$ being the $rms$ of LOFAR surveys; {\\it ii)} we required that the signal from the radio halo is $ \\geq 3 \\times \\xi\\,F$ in at least 5 beam area of LOFAR observations. In both cases we assume a fixed shape of the radial profile of radio halos, calibrated through that of several well studied halos at 1.4 GHz, that introduces a potential source of uncertainty. Although the uncertainties due to the unavoidable simplifications in our calculations, the expected number counts of radio halos highlights the potential of future LOFAR surveys. By assuming the expected sensitivity of the LOFAR all sky survey (\\eg R\\\"ottgering 2009; priv. comm.), rms =0.1 mJy/b, and $\\xi \\sim 2-3$, we predict that about 350 giant radio halos ($\\sim 200$ considering the case {\\it ii)}) can be detected at redshift $\\leq$0.6. This means that LOFAR will increase the statistics of these sources by a factor of $\\sim 20$ with respect to present day surveys. About 55\\% of these halos are predicted with a synchrotron spectral index $\\alpha > 1.9$ between 250-600 MHz, and would brighten only at lower frequencies, unaccessible to present observations. Most important, the spectral properties of the population of radio halos are expected to change with increasing the sensitivity of the surveys as steep spectrum radio halos are expected to populate the low-power end of the RHLF. A large fraction of radio halos with spectrum steeper than $\\alpha \\approx 1.5$ (\\eg Fig.~\\ref{Fig.alpha}) is expected to allow a prompt discrimination between different models for the origin of radio halos, for instance in this case simple energetic arguments would rule out a secondary origin of the emitting electrons (\\eg Brunetti 2004; Brunetti et al. 2008). Due to the large number of expected radio halos, a potential problem with these surveys is the identification of halos and of their hosting clusters. As a matter of fact we expect that LOFAR surveys will open the possibility to unveil radio halos in galaxy clusters with masses $\\gtsim 6-7 \\times 10^{14}$M$_{\\odot}$ at intermediate redshift. On the other hand, statistical samples of X-ray selected clusters, that are unique tools for the identification of the hosting clusters, typically select more massive clusters at intermediate z. Consequently, we explored the potential of the first LOFAR surveys as deep follow ups of available X-ray selected samples of galaxy clusters. \\noindent We calculate the radio halo number counts expected from the follow up of clusters in the eBCS and MACS samples that collect $\\sim 400$ galaxy clusters in the redshift range 0--0.6. We expect that the LOFAR all sky survey, with a planned sensitivity in line with the case $\\xi\\,F=0.25$ mJy/b, will discover about 130 radio halos in eBCS and MACS clusters and that about 40\\% of these radio halos should have very steep spectrum, $\\nu_s \\leq 600$ MHz. The majority of radio halos in eBCS and MACS clusters are expected at z = 0.2-0.4, while the small number of clusters at $z \\geq 0.5$ in the MACS catalog does not allow a statistically solid expectations, although we may expect a couple of radio halos hosted in MACS clusters at this redshift. The MS$^3$ survey will be carried out in 2010, covering the northern hemisphere, and is expected to reach a noise level of about 0.5 mJy/b at 150 MHz, that implies a sensitivity to diffuse emission from galaxy clusters about one order of magnitude (assuming $\\alpha \\approx 1.3$) better than present surveys (\\eg NVSS, Condon et al.~1998; VLSS, Cohen et al.~2007; WENSS, Rengelink et al.~1997). \\noindent We considered MS$^3$ pointings relative to the fields of the about 300 galaxy clusters at $z\\leq0.3$ in the eBCS catalogues. We find that about 60 radio halos are expected to be detected by MS$^3$ observations in these clusters, 25\\% of them (10-15 halos) are expected with $\\nu_s \\leq$ 600 MHz. Fairly sensitive GMRT observations of eBCS clusters at redshift 0.2--0.3 are already available (Venturi et al.~2007, 2008) and we expect that in a few cases radio halos would glow up in the MS$^3$ images where no diffuse radio emission is detected at 610 MHz. We also find that MS$^3$ observations of eBCS clusters at $z=0-0.2$ can be used to test the increase of the fraction of cluster with radio halos with the X-ray luminosity of the host clusters, which is a unique prediction of our model (Fig.\\ref{Fig.histo_Lx}. The most important simplification in our calculations is the use of homogeneous models. Non-homogeneous approaches, that model the spatial dependence of the acceleration efficiency and magnetic field in the halo volume (\\eg Brunetti et al. 2004), and possibly their combination with future numerical simulations, will provide a further step to interpret LOFAR data. Also the use of the extended PS theory is expected to introduce some biases. For instance, it is well-known that the PS mass function underproduces the expected number of massive clusters ($M>10^{15}\\,M_{\\odot}$) at higher redshift, $z\\sim 0.4-0.5$, by a factor of $\\sim 2$ with respect to that found in N-body simulations (\\eg Governato et al. 1999; Bode et al. 2001; Jenkins et al. 2001). Since in our model the great majority of halos at these redshift is associated with massive clusters, the use of the PS mass function implies that the RHNC at $z> 0.4-0.5$ could be underestimated by a similar factor. A further refinement of the approach proposed in the present paper could be obtained with the use of galaxy cluster {\\it merger trees} extracted from N-body simulations. These would also allow for a more {\\it realistic} description of the merger events (spatially resolved, multiple mergers, etc...). In the present paper we focus on a reference set of model parameters. Cassano et al. (2006a) discussed the dependence of model expectations at 1.4 GHz on these parameters. Based on their analysis we expect that all the general results given in the present paper are independent of the adopted values for parameters. The expected number counts of halos should change by a factor of $\\sim2-2.5$ considering sets of model parameters within the region ($$, $b$, $\\eta_t$) that allows for reproducing the observed $P_{1.4}-M_v$ correlation. In this case the number of halos that we expect decreases from super-linear sets of parameters ($b>1$ and $\\geq 1.5\\,\\mu$G) to sub-linear cases ($b<1$ and $\\leq 1.5\\,\\mu$G) (see also Fig.4 in Cassano et al. 2006b); a more detailed study will be presented in a future paper." }, "0910/0910.5928_arXiv.txt": { "abstract": "The radii of some transiting extrasolar giant planets are larger than would be expected by the standard theory. We address this puzzle with the model of coupled radius-orbit tidal evolution developed by \\citet{Ibgui_and_Burrows_2009}. The planetary radius is evolved self-consistently with orbital parameters, under the influence of tidal torques and tidal dissipation in the interior of the planet. A general feature of this model, which we have previously demonstrated in the generic case, is that a possible transient inflation of the planetary radius can temporarily interrupt its standard monotonic shrinking and can lead to the inflated radii that we observe. In particular, a bloated planet with even a circular orbit may still be inflated due to an earlier episode of tidal heating. We have modified our model to include an orbital period dependence of the tidal dissipation factor in the star, $Q'_{\\ast} \\propto P^{\\gamma}$, $-1 \\leqslant \\gamma \\leqslant 1$. With this model, we search, for a tidally heated planet, orbital and radius evolutionary tracks that fall within the observational limits of the radius, the semimajor axis, and the eccentricity of the planet in its current estimated age range. We find that, for some inflated planets (WASP-6b and WASP-15b), there are such tracks; for another (TrES-4), there are none; and for still others (WASP-4b and WASP-12b), there are such tracks, but our model might imply that we are observing the planets at a special time. Finally, we stress that there is a two to three order-of-magnitude timescale uncertainty of the inspiraling phase of the planet into its host star, arising from uncertainties in the tidal dissipation factor in the star $Q'_{\\ast}$. ", "introduction": "\\label{sec:intro} The more than 60 transiting extrasolar giant planets (EGPs) discovered so far\\footnote{ See J. Schneider's Extrasolar Planet Encyclopaedia at http://exoplanet.eu, the Geneva Search Programme at http://exoplanets.eu, and the Carnegie/California compilation at http://exoplanets.org.} offer a unique opportunity to test and improve the models of the structure and evolution of these bodies. The mass and radius of such planets can be inferred from a combination of radial velocity and transit lightcurve measurements that break the planet mass-inclination angle degeneracy. A large theoretical effort has been undertaken for more than a decade now to model and understand the evolution and the radii of transiting planets \\citep{guillot_et_al1996, burrows_et_al2000, Bodenheimer_et_al_2001, burrows_et_al2003, Baraffe_et_al_2003, Bodenheimer_et_al_2003,Gu_et_al_2003, burrows_et_al2004, Fortney_and_Hubbard_2004, Baraffe_et_al_2004, Chabrier_et_al_2004, laughlin_et_al_2005_1, Baraffe_et_al_2005, Baraffe_et_al_2006, burrows_et_al2007, Fortney_et_al_2007, Marley_et_al_2007, chabrier+baraffe2007, Liu_et_al_2008, Baraffe_et_al_2008, Ibgui_and_Burrows_2009, Miller_et_al_2009, Leconte_et_al_2009, Ibgui_et_al_2009_1}. The radius of a gas giant planet depends on many physical effects that are particular to a given planet-star system, including the mass and age of the planet; the stellar irradiation flux and spectrum; the composition -- in particular, the heavy-element content -- of the atmosphere, the envelope, and the core; the atmospheric circulation that couples the day and the night sides; and any processes that could generate an extra power source in the interior of the planet, such as tidal heating. Moreover, the transit radius effect \\citep{burrows_et_al2003,Baraffe_et_al_2003} has to be considered in order to infer the transit radius from the planet's physical radius. Therefore, a custom evolutionary calculation is the most appropriate way to determine a theoretical transit radius, in order to compare it with the observed transit radius. The objective of the present paper is to test the coupled radius-orbit tidal evolution model developed by \\citet{Ibgui_and_Burrows_2009} on some recently discovered inflated planets. We present evolutionary tracks for WASP-4b \\citep{Wilson_et_al_2008, Southworth_et_al_2009_2, Gillon_et_al_2009_1, Winn_et_al_2009_1} and WASP-12b \\citep{Hebb_et_al_2009}. We have also tested the model on TrES-4 \\citep{Mandushev_et_al_2007, Sozzetti_et_al_2008}, WASP-6b \\citep{Gillon_et_al_2009_2}, and WASP-15b \\citep{West_et_al_2009}. Some other inflated planets have been discovered recently, such as HAT-P-13b \\citep{Bakos_et_al_2009_2}, WASP-17b \\citep{Anderson_et_al_2009}, and CoRoT-5b \\citep{Rauer_et_al_2009}. We might apply our formalism to them in the future. The idea of exploring tidal heating as an explanation for the inflated radii was originally formulated by \\citet{Bodenheimer_et_al_2001}. They suggested an excitation mechanism to sustain a nonzero eccentricity, for example a planetary companion \\citep{Bodenheimer_et_al_2003,Mardling_2007}. To date, two transiting EGPs are known to be accompanied by a companion, HAT-P-13b \\citep{Bakos_et_al_2009_2} and HAT-P-7b \\citep{Pal_et_al_2008,Winn_et_al_2009_4}. This reinforces the plausibility of such a scenario. \\citet{Batygin_et_al_2009_2} have coupled a three-body tidal orbital evolution model with a model of the interior structure of HAT-P-13b. They found a quasi-stationary solution and possibly consistent core masses, radii, and tidal heating rates. Assuming, as did \\citet{Liu_et_al_2008}, that the systems are in a quasi-stationary state, \\citet{Ibgui_et_al_2009_1} provide, for each of the systems that they have studied (WASP-6b, WASP-12b, WASP-15b, TrES-4, HAT-P-12b) and for each of the associated atmospheric opacities (solar, $10\\times$solar) a relation between the heavy-element core mass $M_{\\rm core}$ and the ratio $e^2/Q'_{p}$, where $e$ is the orbital eccentricity and $Q'_{p}$ the tidal dissipation factor in the planet. This constraint results from a degeneracy between the dissipation heating rate in the interior of the planet (which increases the radius) and the mass of a possible heavy-element central core (which shrinks the radius). For close-in EGPs (orbital separation $\\sles~0.1-0.15$~AU), tidal torques are strong enough such that they can result in planetary orbital evolution, and they produce tidal heating (dissipation) inside the planet. Such tidal effects were first suggested for transiting EGPs by \\citet{Jackson_et_al_2008_1,Jackson_et_al_2008_2, Jackson_et_al_2008_3}. Jackson et al. included the tides raised on the star and the tides raised on the planet. They found tidal rates close to the levels that \\citet{burrows_et_al2007} proposed to maintain the observed radii of some transiting EGPs. \\citet{Ibgui_and_Burrows_2009} described a model that couples the two consequences of these tidal effects -- planetary radius evolution and orbit evolution. They tested their model on HD~209458b and found an explanation for the radius of this planet. Note that they also showed that a supersolar metallicity of the planetary atmosphere, without invoking tides, can explain the radius. \\citet{Miller_et_al_2009} applied a similar method to all the transiting EGPs, albeit with simplified models for the atmospheres of the planets and parent stars, and a restricted range of possibilities for the tidal dissipation factors $Q'$. We have chosen a more detailed approach by adopting customized atmospheric models and an extended range for $Q'$. Therefore, due to the complexity of the atmospheric calculations, we have selected a couple of planets. We believe that both approaches are complementary. The one followed by \\citet{Miller_et_al_2009} provides a global estimate of the possibilities for matching the observed radii of the transiting EGPs, while our approach, more precise and therefore applied to fewer planetary systems, focuses on more specific issues such as the influence of the atmospheric opacity of the planet, a more detailed model for the tidal dissipation in the star, and a phenomenological study of all the qualitatively possible behaviors of the evolutionary curves \\citep{Ibgui_and_Burrows_2009}. The application of our model to a subset of inflated transiting EGPs is the subject of this paper. The paper outlines, in Section~\\ref{sec:model}, the main assumptions of our coupled radius-orbit tidal evolution model, with a summary of the basic phenomenological results obtained from the previous study by \\citet{Ibgui_and_Burrows_2009}. It also explains our upgraded modeling of the tidal dissipation factor in the star. Section~\\ref{sec:M_Q_a_effect} demonstrates some additional generic results, such as the effect of $M_{\\rm core}$, $Q'_{p}$, and the initial semimajor axis. For each of the following planets, TrES-4, WASP-4b, WASP-6b, WASP-12b, and WASP-15b, we search for evolutionary tracks that fall within the observational limits of the radius, the semimajor axis, and the eccentricity of the planet in its current estimated age range. Our results are described in Section~\\ref{sec:applications}. We have not found any such track for TrES-4. We have found solutions for WASP-6b and WASP-15b. The cases of the planets WASP-4b and WASP-12b, for which we have coupled models that fit, are more interesting. Therefore, we present evolutionary curves for these planets. In fact, the solutions that we obtain for these two planets are valid only for very short age ranges in comparison with the estimated ages of the planets. This would imply that we are observing both planets at a very special time in their evolution, which would be a priori unlikely. In Section~\\ref{sec:plunging_timescale}, we discuss the plunging timescale of a planet into its host star. Its uncertainty can span two to three orders of magnitude. We summarize our results in Section~\\ref{sec:conclusion}.\\\\ ", "conclusions": "\\label{sec:conclusion} We have presented in this paper some new general results of the coupled radius-orbit evolutionary model described in \\citet{Ibgui_and_Burrows_2009}, and we have applied the model to the inflated planets WASP-4b and WASP-12b. We assumed a two-body gravitational and tidal interaction between the planet and its host star, coupling the planetary radius and the orbit evolution. We included the tides raised on the planet and the tides raised on the star. Stellar irradiation and a detailed planetary atmosphere are included. The fundamental result is the transient inflation of the planetry radius that temporarily interrupts its monotonic standard shrinking. An important point is that even though the current orbit of the planet has almost circularized, the radius of the planet can still be inflated due to an earlier episode of tidal heating. This is why we stress that an inflated planet with an observed circular orbit can still have tidal heating as an explanation of its radius. Fixing the planet and star properties, the model is controlled by four free parameters, ($Q'_{p},Q'_{\\ast},e_{i},a_{i}$), that are the tidal dissipation factors in the planet and in the star, and the initial eccentricity and semimajor axis at the beginning of this two-body evolution. We stress the sensitive and nonlinear dependence of the evolutionary curves on these parameters. We have demonstrated that an increase of either the core mass $M_{\\rm core}$, or $Q'_{p}$, or $a_{i}$ results in a lower value of the radius inflation peak and in a delay of its appearance. The final semimajor axis is the same, whatever $M_{\\rm core}$ or $Q'_{p}$, but is larger when $a_{i}$ is larger. At an earlier age, the planet with the larger core has the smaller radius, but this is opposite at later ages. We have enhanced our model by including an orbital period dependence of the tidal dissipation in the star, $Q'_{\\ast} \\propto P^{\\gamma}$, $-1 \\leqslant \\gamma \\leqslant 1$. $Q'_{\\ast}$ drives the inspiral of the planet into its host star. Applications of our model to recently detected transiting inflated planets show that: \\begin{itemize}\\itemsep-0.04in% \\vspace{-6pt} \\item WASP-6b and WASP-15b can be fit at solar opacity over Gyr age ranges. \\item We have not found an acceptable fit for TrES-4, at either solar, $3\\times$solar, or $10\\times$solar planet atmospheric opacity. \\item WASP-4b can be fit at solar opacity with, for example, the combination ($Q'_{p},e_{i},a_{i}$)=($10^{8.0},0.80,0.050$) and with $Q'_{\\ast}=10^{6.5}\\times(P/\\rm 10days)^{-1}$. \\item WASP-12b can be fit at solar opacity with, for example, the combination ($Q'_{p},e_{i},a_{i}$)=($10^{8.0},0.73,0.055$) and with $Q'_{\\ast}=10^{6.5}\\times(P/\\rm 10days)^{-1}$. \\end{itemize} \\vspace{-6pt} For WASP-4b and WASP-12b, the ranges of ages that allow simultaneous fits of radius, semimajor axis, and orbital eccentricity, are very narrow, seeming to suggest that, if the two-body coupled evolutionary model described herein is in fact responsible for these planets' inflated radii, then we are observing them at a special epoch in their evolution. However, relaxing the fit-criterion from 1$\\sigma$ to 2$\\sigma$ or 3$\\sigma$ would alleviate this apparent problem. Our results (in particular, for TrES-4) suggest that a coupled radius-orbit tidal evolution model might not on its own explain the radii of all the inflated transiting giant planets. An alternative scenario with stationary heating has been proposed \\citep{Ibgui_et_al_2009_1} and applied to all the planets discussed in this paper. Though not providing direct solutions to the inflated radii issue, this scenario constrains the ratio $e^{2}/Q'_{p}$ for a given $M_{\\rm core}$. Finally, a combination of these two models could be imagined with a two-body interaction, followed by a quasi-steady low eccentricity phase due to perturbations by a second planet. The last point we make in this paper is the uncertainty of the plunging timescale during the spiraling of the planet into its host star. This timescale is strongly dependent on the semimajor axis; specifically, it depends on $a$ to the 6.5 power. It also has a linear dependence on $Q'_{\\ast}$, which is a parameter that is uncertain by several orders of magnitude. We have shown that HD~209458b can plunge in between 0.5 and 60~Gyr from now, a 2-order-of-magnitude range, and that WASP-12b can plunge in between 0.1 and 100~Myr from now, a 3-order-of-magnitude range. \\citet{Ibgui_and_Burrows_2009} have suggested caveats to, and ways to improve, the model employed here. We close by noting several additional points. The orbital evolution equations depend on the theory of tidal dissipation inside gaseous planets and stars. Improvements to this theory might result in different evolutionary tracks (of $R_p$, $e$, and $a$) from the ones presented in this paper. Furthermore, we have noted that it is not strictly appropriate to apply these equations to model scenarios with large values of orbital eccentricity. However, both for comparison with previous work and because the proper tidal theory remains unknown, our present approach is a valuable step in exploring the extent to which tidal dissipation might explain the radii of the inflated EGPs. Further observations that might help to constrain this model and to discriminate between this and the stationary-state model of \\citet{Ibgui_et_al_2009_1} include both increasing the accuracy of orbital eccentricity measurements and searching for companions to the transiting EGPs. These and other advances will help us progress toward a better understanding of the puzzle of the inflated planets." }, "0910/0910.0399_arXiv.txt": { "abstract": "Sagittarius\\,A$^\\star$ (\\sgra) is the supermassive black hole residing at the center of the Milky Way. It has been the main target of an extensive multiwavelength campaign we carried out in April 2007. Herein, we report the detection of a bright flare from the vicinity of the horizon, observed simultaneously in X-rays (\\xmm/EPIC) and near infrared (\\vlt/NACO) on April 4$^{\\rm th}$ for 1--2~h. For the first time, such an event also benefitted from a soft $\\gamma$-rays (\\integ/ISGRI) and mid infrared (\\vlt/VISIR) coverage, which enabled us to derive upper limits at both ends of the flare spectral energy distribution (SED). We discuss the physical implications of the contemporaneous light curves as well as the SED, in terms of synchrotron, synchrotron self-Compton and external Compton emission processes. ", "introduction": "From the discovery of a compact radio source, \\sgra, at the Galactic Center (GC) in 1974 \\citep{balick74} to the near infrared (NIR) tracking of stars in Keplerian motion around \\sgra~three decades later \\citep{schodel02,ghez03}, the evidence for a $\\sim$$4\\times10^{6}$~$M_\\odot$ black hole with very slow proper motion at the dynamical center of our galaxy \\citep{reid08} gradually piled up \\citep[see][for a general review and references therein]{melia07}. Yet, the long quest for the high energy emission pertaining to the black hole has only been achieved recently. \\sgra~was resolved as a notably dim ($2.4\\times10^{33}$ \\ergpersec, 2--10~keV) and slightly extended (1.4$''$) point source with the \\chandra~satellite in 1999 \\citep{baganoff03a}. One year later, the same instrument witnessed the source exhibiting an X-ray flare for $\\sim$3 h \\citep{baganoff01}. A $\\sim$10~min long substructure within the light curve of the eruption and light time travel arguments imply that this event took place close to the event horizon ($<15~R_{\\rm S}$). Many other detections of X-ray flares followed, either with \\xmm~or \\chandra~\\citep[see e.g.][]{goldwurm03a,baganoff03b,porquet03,belanger05}, and established that the duty cycle of the black hole is nearly one X-ray flare per day. The origin of these events is still unclear, in spite of all the efforts aimed at their monitoring in different energy ranges. In 2003, NIR flares from \\sgra~were indeed discovered with the \\vlt~\\citep{genzel03}, and later confirmed by the \\keck~\\citep{ghez04} and the \\hst~\\citep{yusef-zadeh06a}. They occur more frequently than the X-ray ones (around four per day) and have been observed in many NIR atmospheric pass bands (H, K, L, M). Each new infrared flare has generally induced either spatial \\citep{clenet05}, spectral \\citep{eisenhauer05,ghez05,gillessen06,krabbe06,hornstein07}, polarimetric \\citep{eckart06b,meyer06,meyer07,trippe07}, or timing studies \\citep{meyer08,do09}. Numerous multiwavelength campaigns showed that an X-ray flare always comes along with a simultaneous NIR one\\footnote{The converse is not true, some NIR flares have no X-ray counterpart \\citep{hornstein07}.} \\citep{eckart04,eckart06a,eckart08a,yusef-zadeh06a,hornstein07}, and maybe a delayed submm one \\citep{marrone08,eckart08b,yusef-zadeh08} caused by plasmon expansion \\citep{liu04,yusef-zadeh06b}. Above 6~$\\mu$m, in the mid infrared (MIR), no detection of \\sgra~has been reported so far. Recent upper limits on the black hole flux at 8.6~$\\mu$m were set by the \\vlt/VISIR instrument during low level NIR variability by \\citet{schodel07}, who argued that a detection would be reachable in case of a strong NIR flare. Above 20 keV, repeated surveys of the heart of the Milky Way in soft $\\gamma$-rays with the \\integ~satellite unveiled a persistent pointlike source compatible with \\sgra~location (within the 1$'$ error radius), \\igr~\\citep{belanger04,belanger06}. The nature of the source is still uncertain, and a possible association with the supermassive black hole remains conceivable. Given the limited angular resolution of the soft $\\gamma$-ray telescope \\integ/IBIS/ISGRI ($\\sim$12$'$ FWHM), the best way to unequivocally identify the mysterious \\igr~with \\sgra~is the detection of correlated variability between soft $\\gamma$-rays and other wavelengths. To tackle the above puzzles and investigate the correlated X-ray/NIR variability of \\sgra~in more details, a coordinated multiwavelength campaign on the GC was conducted in spring 2007. It involved in particular the \\xmm~and \\integ~satellites for the high energies, as well as the \\vlt/ NACO and \\vlt/VISIR ground instruments to cover the NIR and MIR part of the spectrum, respectively. Their results are presented in Sect.~\\ref{obs} and interpreted in Sect.~\\ref{SED}. Note that the X-ray and infrared findings have already been published by \\citet{porquet08} and \\citet{dodds09}, respectively. We will not discuss here the short term variability of \\sgra~in April 2007 at cm, mm, and submm wavelengths, which will be reported in another article, along with NIR results obtained by the Hubble Space Telescope \\citep{yusef-zadeh09}\\footnote{For further discussion of the past variability of \\sgra~in cm and mm bands, see for example \\citet{zhao01,herrnstein04} and \\citet{tsuboi99,zhao03}, respectively.}. Throughout this paper we adopt a GC distance of 8 kpc \\citep{reid93} and a black hole mass $M_{\\bullet}=4\\times10^{6}$~$M_\\odot$ \\citep{ghez08}, for which the Schwarzschild radius is $R_{\\rm S}=1.2\\times10^{12}$~cm. \\begin{figure*} \\centering \\includegraphics[trim=1cm 2cm 1cm 1.5cm, clip=true, width=14cm]{Fig1.pdf} \\caption{Multiwavelength and multiscale views of the Galactic Center in April 2007 in R.A. ($^\\circ$) horizontally and Dec. ($^\\circ$) vertically (North and East point towards the top and the left, respectively). \\emph{Top, left}: INTEGRAL/ISGRI mosaic in the 20--40~keV band. The Galactic plane runs from upper left to bottom right. \\emph{Top, right}: INTEGRAL/JEM-X 1 mosaic in the 3--20~keV band. \\emph{Middle}: XMM/PN image in the 2--10~keV band. \\emph{Bottom, left}: VLT/NACO image at the peak of the flaring period at 3.8~$\\mu$m. \\emph{Bottom, right}: VLT/VISIR average image of the 3$^{\\rm rd}$/4$^{\\rm th}$ April night at 11.88~$\\mu$m. The three dusty arms swirling around \\sgra~compose the so-called ``mini-spiral''.} \\label{ima} \\end{figure*} ", "conclusions": "This paper complements a series of articles about the April 2007 synchronous observations of the Galactic Center from radio to $\\gamma$-rays \\citep{porquet08,dodds09,yusef-zadeh09}. Here, we have recapped the results on the brightest flare ever detected simultaneously at NIR and X-ray frequencies. We have also reported for the first time $\\gamma$-ray constraints on such an event, which, added to our MIR/NIR/X-ray spectral measurements, constitute the broadest simulaneous spectrum of a flare ever achieved. The essential observational conclusions may be summarized as follows: \\begin{itemize} \\item the peaks of the X-ray and NIR emissions are coincident within 3~min; \\item the width of the NIR flare light curve is broader than the X-ray one by a factor $\\sim$2; \\item the NIR light curve is substructured on a timescale of $\\sim$20~min while the X-ray light curve is rather smooth; \\item there is no detectable MIR counterpart; \\item the soft $\\gamma$-ray source \\igr~is non variable. \\end{itemize} The high quality of the spectral information we gathered allowed for a discussion of the several classical radiative processes models employed to explain the flares: SSC, EC, and SB. Yet, none of these mechanisms is entirely satisfactory to meet our observations. The theoretical inquiries to come will have to take into account the time evolution of the phenomenon and the aging of the radiating particles to better connect the light curves and spectra. From an observational stand point, it will be useful to repeat such NIR/X-ray measurements in a near future to get two respective individual and fully contemporaneous spectra, which has never been accomplished thus far. As we have seen, one key probe of what powers the flares, is a better determination of the X-ray spectral slope. In a more distant future, thanks to a broad X-ray sensitivity over the 1--80~keV band and a high angular resolution above 10~keV, \\simx~should address this issue and resolve the GC region in soft $\\gamma$-rays." }, "0910/0910.5082_arXiv.txt": { "abstract": "{ The recently discovered planetary system HD45364 which consists of a Jupiter and Saturn mass planet is very likely in a 3:2 mean motion resonance. The standard scenario to form planetary commensurabilities is convergent migration of two planets embedded in a protoplanetary disc. When the planets are initially separated by a period ratio larger than two, convergent migration will most likely lead to a very stable 2:1 resonance for moderate migration rates. To avoid this fate, formation of the planets close enough to prevent this resonance may be proposed. However, such a simultaneous formation of the planets within a small annulus, seems to be very unlikely. Rapid type III migration of the outer planet crossing the 2:1 resonance is one possible way around this problem. In this paper, we investigate this idea in detail. We present an estimate for the required convergent migration rate and confirm this with N-body and hydrodynamical simulations. If the dynamical history of the planetary system had a phase of rapid inward migration that forms a resonant configuration, we predict that the orbital parameters of the two planets are always very similar and hence should show evidence of that. We use the orbital parameters from our simulation to calculate a radial velocity curve and compare it to observations. Our model can explain the observational data as good as the previously reported fit. The eccentricities of both planets are considerably smaller and the libration pattern is different. Within a few years, it will be possible to observe the planet-planet interaction directly and thus distinguish between these different dynamical states. } \\date{Submitted: 30 August 2009 - Revised: 26 October 2009 - Accepted: 26 October 2009 } ", "introduction": "Over 400 extrasolar planets have already been discovered \\citep{exoplanet} and their diversity keeps challenging planet formation theory. For example, the recently discovered multi-planetary system HD45364 raises interesting questions about its formation history. The planets have masses of $m_1=0.1906M_{\\text{Jup}}$ and $m_2=0.6891M_{\\text{Jup}}$ and are orbiting the star at a distance of $a_1=0.6813~\\text{AU}$ and $a_2=0.8972~\\text{AU}$, respectively \\citep{CorreiaUdry2008}. The period ratio is close to $1.5$ and a stability analysis implies that the planets are deep inside a 3:2 mean motion resonance. The planets have most likely formed further out in cooler regions of the proto-stellar disc, as water ice which is an important ingredient for dust aggregation can only exist beyond the snow line which is generally assumed to be at radii larger than $2~\\text{AU}$ \\citep{SasselovLecar2000}. It is then usually assumed that migration due to planet disc interactions has moved the planets closer to the star. Although the details of this process are still hotly debated, the existence of many resonant multiplanetary systems supports this idea. During migration the planets can get locked into a commensurabilities after which the planets migrate together with a constant period ratio. In such a resonance, one or more resonant angles are librating \\citep[see e.g.][]{LeePeale01}. For the planetary system HD45364 this standard picture poses a new problem. Assuming that the planets have formed far apart from each other, the outcome for the observed masses after migration is almost always a 2:1 mean motion resonance, not 3:2 as observed. The 2:1 resonance that forms is found to be extremely stable. One possible way around this is a very rapid convergent migration phase that passes quickly through the 2:1 resonance. In this paper we explore this idea quantitatively. The plan of the paper is as follows. In section \\ref{sec:formationnbody} we show, using N-body simulations, that the system always ends up in the 2:1 resonance assuming moderate migration rates. We present scenarios which result in formation of a 3:2 resonance after a rapid migration phase. In section \\ref{sec:formationhydro} we perform hydrodynamic simulations with a variety of disk models in order to explore the dependence on the physical setup. In section \\ref{sec:otherformationscenarios} we briefly discuss other formation scenarios. We go on to compare the orbital parameters of our simulations with the observed radial velocity data in section \\ref{sec:observation}. We find that the orbital parameters observed in our simulations differ from those estimated from a statistical fit by \\cite{CorreiaUdry2008}. However, our models reproduce the observational data very well and we argue that our fit has the same level of significance as that obtained by \\cite{CorreiaUdry2008}. Future observations will be able to resolve this issue. This is the first prediction of orbital parameters for a specific extrasolar planetary system derived from planet migration theory only. The parameter space of orbital configurations produced by planet disc interactions (low eccentricities, relatively small libration amplitudes) is very small. As in case of the GJ876 system, this can provide strong evidence on how the system formed. Finally, we summarise our results in section \\ref{sec:conclusion}. ", "conclusions": "The planets in the multi-planetary system HD45364 are most likely in a 3:2 mean motion resonance. This poses interesting questions on it's formation history. Assuming that the planets form far apart from each other and migrate with moderate migration rate, as predicted by standard planet formation and migration theories, the most likely outcome is a 2:1 mean motion resonance, contrary to the observation of a 3:2 MMR. In this work, we investigated a possible way around this problem by letting the outer planet undergo a rapid inward type~III migration. We presented an analytical estimate and performed N-body as well as hydrodynamical simulations. We found that it is indeed possible to form a 3:2 MMR and avoid the 2:1 resonance, thus resembling the observed planetary system using reasonable disc parameters. Hydrodynamical simulations suggest that the system is more likely to sustain the resonance for large aspect ratios, as the migration of the inner planet is slowed down, thus avoiding divergent migration. Finally, we used the orbital configuration found in the hydrodynamical formation scenario to calculate a radial velocity curve. This curve has then been compared to observations and the resulting fit has an identical $\\chi^2$ value than the previously reported \\textit{best fit}. Our solution is stable for at least a million years. It is in a dynamically different state, both planets having lower eccentricities and a different libration pattern. This is the first time that planet migration theory can predict a precise orbital configuration of a multiplanetary system. This might also be the first direct evidence for type III migration if this scenario turns out to be true. The system HD45364 remains an interesting object for observers, as the differences between the two solutions can be measured in radial velocity within a couple of years. \\label{sec:conclusion}" }, "0910/0910.0458_arXiv.txt": { "abstract": "The timescale for energy release is an important parameter for constraining the coronal heating mechanism. Observations of ``warm'' coronal loops ($\\sim 1$\\,MK) have indicated that the heating is impulsive and that coronal plasma is far from equilibrium. In contrast, observations at higher temperatures ($\\sim 3$\\,MK) have generally been consistent with steady heating models. Previous observations, however, have not been able to exclude the possibility that the high temperature loops are actually composed of many small scale threads that are in various stages of heating and cooling and only appear to be in equilibrium. With new observations from the EUV Imaging Spectrometer (EIS) and X-ray Telescope (XRT) on \\textit{Hinode} we have the ability to investigate the properties of high temperature coronal plasma in extraordinary detail. We examine the emission in the core of an active region and find three independent lines of evidence for steady heating. We find that the emission observed in XRT is generally steady for hours, with a fluctuation level of approximately 15\\% in an individual pixel. Short-lived impulsive heating events are observed, but they appear to be unrelated to the steady emission that dominates the active region. Furthermore, we find no evidence for warm emission that is spatially correlated with the hot emission, as would be expected if the high temperature loops are the result of impulsive heating. Finally, we also find that intensities in the ``moss,'' the footpoints of high temperature loops, are consistent with steady heating models provided that we account for the local expansion of the loop from the base of the transition region to the corona. In combination, these results provide strong evidence that the heating in the core of an active region is effectively steady, that is, the time between heating events is short relative to the relevant radiative and conductive cooling times. ", "introduction": "The coronal heating problem is one of the most fundamental open questions in solar physics. The formation of the high temperature solar atmosphere is clearly related to the magnetic fields generated in the solar interior, but how this magnetic energy is converted to the thermal energy of the corona remains unknown. One important constraint on the coronal heating mechanism is the time scale for energy release. If energy is deposited into magnetic flux tubes on timescales that are very short compared to a characteristic cooling time, then the corona will be filled with loops that are close to equilibrium and appear to be steady. Active region observations have generally suggested that high temperature loops are consistent with steady heating \\citep[e.g.,][]{porter1995,kano1995}. These studies have found that the observed evolution of the emission is much slower than the radiative and conductive cooling timescales. There has also been some success in modeling entire active regions with steady heating models \\citep{schrijver2004,warren2006b,winebarger2008,lundquist2008}. Finally, \\cite{antiochos2003} have argued that observations of the ``moss,'' the footpoints of high temperature loops, are also consistent with steady heating. They found that the average moss intensities are typically constant over many hours and loops cooling through 1\\,MK were not observed in the moss region they observed. There is considerable evidence that coronal loops observed at lower temperatures ($\\sim 1$\\,MK) are evolving and not in equilibrium \\citep[e.g.,][]{aschwanden2001b,winebarger2003b,cirtain2007,urra2009}. These ``warm'' loops appear to have apex densities that are much higher than can be accounted for by steady heating models \\citep[e.g.,][]{winebarger2003,aschwanden2008b}. The properties of the warm loops are more consistent with impulsive heating models \\citep[e.g.,][]{warren2002b,spadaro2003,warren2003}. If it is true that hot loops are close to equilibrium while the warm loops are generally heated impulsively then the coronal heating mechanism becomes even more difficult to understand. The cooling time for short, hot loops is relatively rapid (typically a few hundred seconds), so heating events on these loops would need to occur very frequently. The cooling time for the warm loops is much longer (typically a few thousand seconds), so impulsive heating events on these loops would be infrequent. It is tempting to conjecture that the emission at high temperatures is also consistent with impulsive heating models and that the apparent steadiness of this emission is the result of the superposition of many evolving stands along the line of sight \\citep[e.g.][]{cargill1997,cargill2004,patsourakos2006}. Previous observations have not been able to exclude this possibility. For example, it could be that the cooling loops are not easy to detect in an active region core because they are faint relative to both the bright moss and the extended corona in which the active region is embedded. The EUV Imaging Spectrometer (EIS) and the X-ray Telescope (XRT) on the \\textit{Hinode} mission provide a new opportunity to observe active region emission in unprecedented detail. EIS observes emission covering a very broad range of coronal temperatures: between \\ion{Fe}{7} and \\ion{Fe}{24}, only 4 ionization stages of Fe are not present in the EIS data (\\ion{Fe}{18}--\\textsc{xxi}). EIS also observes \\ion{Ca}{14}, \\textsc{xv}, \\textsc{xvi}, and \\textsc{xvii}, providing excellent coverage of the critical temperature range around 3\\,MK \\citep{warren2008c}. XRT is a broadband imaging telescope that observes high temperature plasma very efficiently and at high spatial resolution. The high cadence XRT data complement EIS, which often observes at a much lower cadence, and allows us to track the evolution of coronal plasma over a large field of view. In this paper we use EIS and XRT observations to examine the properties of coronal plasma in the core of an active region. The active region that we have selected (NOAA active region 10960) is unusual in that most of the overlying warm loops are located to the north and south of the region, providing a largely unobstructed view of the active region core over a wide range of temperatures (see Figure~\\ref{fig:context}). High cadence EIS observations of this region, which was observed by \\textit{Hinode} during the period 4--13 June 2007, have shown that the \\ion{Fe}{12} 195.119\\,\\AA\\ intensities, Doppler shifts, and non-thermal widths in the moss are constant over long periods of time, suggesting steady heating \\citep{brooks2009b}. In this study we find three additional lines of evidence that also indicate that the heating of high temperature loops is steady. We examine the evolution of the emission in individual pixels in XRT and find that the vast majority of the emission is constant to within approximately 15\\% during this period. The inspection of emission at different temperatures in the core of the active region shows that there is no evidence for loops cooling from high temperatures. We find no relationship between the warm emission (\\ion{Fe}{10}--\\ion{Fe}{14}) and the steady hot emission (\\ion{Ca}{14}--\\ion{Ca}{17}). Finally, we find that we can bring all of the observed moss intensities into agreement with steady heating models, if we allow for loop constriction at the base of the loop. Previous modeling work had found significant discrepancies at the lowest temperatures observed with EIS \\citep{warren2008}. None of these observations is conclusive; they do not necessarily exclude alternative models that involve non-equilibrium processes. However, these observations do provide very strong constraints on the mechanism responsible for producing the high temperature emission observed in solar active regions. ", "conclusions": "We have presented a comprehensive analysis of observations in the core of an active region using data from the EIS and XRT instruments on \\textit{Hinode} and \\textit{TRACE}. The apparent steadiness of the XRT emission, the lack of spatial correlation between the hot and warm emission, and the consistency of the funnel models with the observed emission all point to frequent heating events that keep the hot loops close to equilibrium. Furthermore, these results are consistent with high cadence EIS measurements of moss intensities, Doppler shifts, and nonthermal widths that show little evidence of dynamical events over many hours \\citep{brooks2009b}. In combination, these results provide strong evidence that the heating in the core of an active region is effectively steady, that is, the time between heating events is short relative to the relevant radiative and conductive cooling times." }, "0910/0910.2946_arXiv.txt": { "abstract": "A modern Q-band low noise amplifier (LNA) front-end is being fitted to the 10.4~m millimeter-wave telescope at the Raman Research Institute (RRI) to support observations in the 40-50~GHz frequency range. To assess the suitability of the surface for this purpose, we measured the deviations of the primary surface from an ideal paraboloid using radio holography. We used the 11.6996 GHz beacon signal from the GSAT3 satellite, a 1.2~m reference antenna, commercial K$_{u}$-band Low Noise Block Convereters (LNBC) as the receiver front-ends and a Stanford Research Systems (SRS) lock-in amplifier as the backend. The LNBCs had independent free-running first local oscillators (LO). Yet, we recovered the correlation by using a radiatively injected common tone that served as the second local oscillator. With this setup, we mapped the surface deviations on a $64 \\times 64$ grid and measured an rms surface deviation of $\\sim 350~\\mu$m with a measurement accuracy of $\\sim 50~\\mu$m. ", "introduction": "RRI has a mm-wave Leighton telescope, of 10.4~m diameter with 81 hexagonal panels~\\citep{ref:tks}. It is being rejuvenated to undertake Q-band observations at 43~GHz. This requires the surface rms error to be below $\\sim 370~\\mu$m in order to have an aperture efficieny better than 50\\%, as given by Ruze's relation:~$\\eta_{ap}=\\eta_0 ~exp{(\\frac{4\\pi \\delta}{\\lambda})^2}$~\\citep{ref:ruze66}. Radio holographic surface measurement was carried out in Aug-Sep 2007 to measure the surface rms error, and to identify panels that may need correction. In this poster paper, we report the details of this experiment, analyse the data, discuss the results and present our conclusions. \\begin{figure}[!t] \\centering \\epsfig{figure=holoRx4.eps, width=0.93\\linewidth} \\caption{\\small Dual channel holography receiver layout\\normalsize } \\label{f:receiver} \\end{figure} ", "conclusions": "The effect of free runnning local oscillators in LNBC was solved by radiatively injecting a tone. The effect of satellite drifting was removed by pointing to the satellite every 30~minutes. An SNR $>$ 5000 was achieved by making the receiver chain robust, repeating and co-adding the central block scans \\& using satellite pointing data for amplitude and phase calibration. From the measured surface deviations (see Fig.~\\ref{f:map}(a)), the surface rms error is calculated to be $\\sim350~\\mu$m. It is within $\\lambda$/16, implying that Q Band observations are possible. It can also be seen that some panels require correction. Fig.~\\ref{f:map}(b) shows the residual obtained by subtracting two independent surface deviation measurements. The rms of this residual is $\\sim70~\\mu$m. Therefore, the measurement accuracy is estimated to be $\\sim50~\\mu$m." }, "0910/0910.3037_arXiv.txt": { "abstract": "Based on the data obtained from the \\textit{Spitzer}/GLIPMSE Legacy Program and the 2MASS project, we derive the extinction in the four IRAC bands, [3.6], [4.5], [5.8] and [8.0]$\\mum$, relative to the 2MASS $\\Ks$ band (at 2.16$\\mum$) for 131 GLIPMSE fields along the Galactic plane within $|l|\\leq65^{\\rm o}$, using red giants and red clump giants as tracers. As a whole, the mean extinction in the IRAC bands (normalized to the 2MASS $\\Ks$ band), $A_{[3.6]}/A_\\Ks\\approx0.63\\pm0.01$, $A_{[4.5]}/A_\\Ks\\approx0.57\\pm0.03$, $A_{[5.8]}/A_\\Ks\\approx0.49\\pm0.03$, $A_{[8.0]}/A_\\Ks\\approx0.55\\pm0.03$, exhibits little variation with wavelength (i.e. the extinction is somewhat flat or gray). This is consistent with previous studies and agrees with that predicted from the standard interstellar grain model for $R_V=5.5$ by \\citet{Weingartner01}. As far as individual sightline is concerned, however, the wavelength dependence of the mid-infrared interstellar extinction $A_{\\lambda}/A_\\Ks$ varies from one sightline to another, suggesting that there may not exist a ``universal'' IR extinction law. We, for the first time, demonstrate the existence of systematic variations of extinction with Galactic longitude which appears to correlate with the locations of spiral arms as well as with the variation of the far infrared luminosity of interstellar dust. ", "introduction": "With the development of space infrared (IR) astronomy, the precise determination of IR extinction becomes urgent in order to recover the intrinsic colors and spectral energy distributions (SEDs) of heavily obscured sources. There have been various attempts to measure the IR extinction based on the \\emph{Infrared Space Observatory}(\\emph{ISO}) and \\emph{Spitzer Space Telescope} since \\citet{Lutz96} obtained the mid-IR extinction from several hydrogen recombination lines and demonstrated the absence of the model-predicted pronounced minimum around 7$\\mum$. This was supported by \\citet{Jiang03, Jiang06} based on the ISOGAL database \\citep{Omont99}, and by \\citet{Indebetouw05} based on the data from the \\emph{Spitzer} Galactic Legacy Infrared Midplane Survey Extraordinaire (GLIMPSE) Legacy Program \\citep{Benjamin03}. All these results roughly agree with the extinction predicted by the standard interstellar grain model for $R_V=5.5$ of \\citet{Weingartner01} and \\citet{Draine03}.\\footnote{% $R_V\\equiv A_V/E(B-V)$ is the total-to-selective extinction ratio, where $E(B-V)\\equiv A_B-A_V$, the color excess, is the difference between the extinction in $B$ and $V$ bands. } However, so far only a few wave bands have been investigated and the sky coverage is also limited. No consensus has been reached yet regarding the interstellar extinction at $\\simali$5--8$\\mum$. Recent progress in the IR extinction measurements is made toward star-forming regions mainly based on the \\emph{Spitzer} observations. Thanks to the high sensitivity of \\emph{Spitzer}, deep photometry is now possible and objects that suffer severe extinction are now reachable. \\citet{Flaherty07} studied five nearby star-forming regions at mid-IR wavelengths (3.6$\\mum$, 4.5$\\mum$, 5.8$\\mum$ and 8.0$\\mum$ from the InfraRed Array Camera [IRAC],\\footnote{% The effective wavelengths of the four IRAC bands are actually 3.545$\\mum$, 4.442$\\mum$, 5.675$\\mum$ and 7.760$\\mum$, respectively. } and 24$\\mum$ from the Multiband Imaging Photometer [MIPS]). They confirmed a relatively flat extinction curve at $\\simali$4--8$\\mum$. \\citet{Roman07} studied a star-forming dense cloud core located in the Pipe Nebula, and found that the IR extinction in the IRAC bands of that region also agrees with the $R_{V}=5.5$ model curve and indicates a dust size distribution favoring larger sizes. Although both the \\emph{ISO} and \\emph{Spitzer} measurements agree with each other in that the $\\simali$5--8$\\mum$ extinction is relatively flat and lacks the model-predicted minimum around 7$\\mum$ of \\citet{Draine89}, there do exist differences among various measurements made for different sightlines. In their sample of five sightlines toward star-forming regions, \\citet{Flaherty07} found a clear difference between one sightline and the other four sightlines (see their Table 3). They derived higher $A_{\\lambda}/A_\\Ks$ ratios and a flatter wavelength dependence than that of \\citet{Indebetouw05} for the same sightline toward $l=284^{\\circ}$ in the Galactic Plane. From $>$\\,200 fields observed in the ISOGAL survey, \\citet{Jiang06} analyzed the extinction at 7$\\mum$ and 15$\\mum$ along $\\simali$120 directions. They found marginal variation of the extinction at 7$\\mum$. It is commonly believed that, with the parameter $R_V$ increasing in denser regions, the variation of the ultraviolet (UV) and visual extinction with wavelength becomes flatter than that of the diffuse interstellar medium (ISM) which is characterized with a lower $R_V$ \\citep{Cardelli89}, while the near-IR extinction seems to be ``universal\", with little variation among different sightlines \\citep{Draine89}. However, \\citet{Nishiyama06a, Nishiyama09} recently argued against such a universal near-IR extinction. In addition, \\citet{Fitzpatrick04} argued that the IR-through-UV Galactic extinction curves should not be considered as a simple one-parameter family, whether characterized by $R_V$ as suggested by \\citet{Cardelli89} or any other parameters. \\citet{Whittet77} presented observational evidence for a small but appreciable variation in $R_V$ with Galactic longitude. He suggested that the most likely explanation for this is a variation in the mean size of the dust in the local spiral arm. However, unfortunately only a few data points were used in that work and therefore no systematic variation of the extinction with Galactic longitude was reported. \\citet{Jiang06} obtained the extinction around 7$\\mum$ for 129 different sightlines and no clear variation with Galactic longitude was found, although the extinction ratio $A_{[7]}/A_\\Ks$ does appear to exhibit a tendency of decreasing toward the Galactic center where $|l|<2^{\\circ}$\\citep{Jiang06}. The GLIMPSE Legacy Program surveyed the Galactic plane, with a large area coverage ($|l|\\leq65^{\\circ}$) and a detection limit of $\\simali$15.5--13.0$\\magni$ from 3.6 to 8.0$\\mum$ \\citep{GQA}. It provides an opportunity to explore the systematic variation of interstellar extinction in the IR with Galactic longitude. In this work, we explore whether the mid-IR extinction varies among sightlines and how it varies in different interstellar environments based on the \\emph{Spitzer}/GLIMPSE database. In \\S\\ref{DATA}, the GLIMPSE data used in this work is briefly described. \\S\\ref{method} presents the method adopted to derive the extinction. In \\S\\ref{tracer} we discuss the selection of two different types of tracers (i.e. red giants and red clump giants). \\S\\ref{results} reports the resulting extinction ratios $A_{\\lambda}/A_\\Ks$ and the mean extinction from the total 131 GLIMPSE fields. Also discussed in \\S\\ref{results} are the comparison of the extinction derived here with previous studies performed by \\citet{Indebetouw05} and \\citet{Flaherty07}, and the longitudinal variation of the extinction ratios $A_{\\lambda}/A_\\Ks$ as well as its relation with the Galactic spiral arms and the distribution of interstellar dust. In \\S6 we summarize our major conclusions. ", "conclusions": "Using red giants and red clump giants as tracers, we have derived $A_\\lambda/A_\\Ks$, the extinction (relative to the 2MASS $\\Ks$ band) in the four IRAC bands, [3.6], [4.5], [5.8] and [8.0]$\\mum$ for 131 GLIPMSE fields along the Galactic plane within $|l|\\leq65^{\\rm o}$, based on the data obtained from the \\textit{Spitzer}/GLIPMSE Legacy Program and the 2MASS Survey project. The principal results of this paper are the following: \\begin{enumerate} \\item The mean extinction in the IRAC bands (normalized to the 2MASS $\\Ks$ band), $A_{[3.6]}/A_\\Ks\\approx0.63\\pm0.01$, $A_{[4.5]}/A_\\Ks\\approx0.57\\pm0.03$, $A_{[5.8]}/A_\\Ks\\approx0.49\\pm0.01$, and $A_{[8.0]}/A_\\Ks\\approx0.55\\pm0.03$, exhibits little variation with wavelength and lacks the minimum at $\\simali$7$\\mum$ predicted from the standard interstellar grain model for $R_V=3.1$. This is consistent within errors with previous observational determinations based on {\\it ISO} and {\\it Spitzer} data and with that predicted from the grain model for $R_V=5.5$ of \\citet{Weingartner01}. \\item The wavelength dependence of interstellar extinction in the mid-IR varies from one sightline to another, suggesting that there may not exist a ``universal'' IR extinction law. \\item There exist systematic variations of extinction with Galactic longitude which appears to correlate with the locations of spiral arms and with the variation of the 240$\\mum$ dust emission. This can be understood in terms of larger grain sizes (arising from coagulational growth), enhanced dust concentration, and higher starlight intensities in the spiral arm regions. \\end{enumerate}" }, "0910/0910.3989_arXiv.txt": { "abstract": "Extrasolar planets are not named and are referred to only by their assigned scientific designation. The reason given by the IAU to not name the planets is that it is considered impractical as planets are expected to be common. I advance some reasons as to why this logic is flawed, and suggest names for the 403 extrasolar planet candidates known as of Oct 2009. The names follow a scheme of association with the constellation that the host star pertains to, and therefore are mostly drawn from Roman-Greek mythology. Other mythologies may also be used given that a suitable association is established. ", "introduction": "Since the discovery of the first extrasolar planet, around the star 51 Pegasi (Mayor \\& Queloz 1995), over 400 planets surrounding other stars have been discovered. It is no exaggeration to say that, for astronomy, the year of 1995 has a historic resonance with 1781, 1846, and 1930. However, unlike Uranus, Neptune, and Pluto, the almost totality of these extrasolar planets are known by no other name than the scientifically dry designations given to them. It is my intent to make the case that naming these planets is desirable. Poincar\\'e (1905) emphasized the usefulness of astronomy by saying that ``it is useful because it raises us above ourselves, because it is great, because it is beautiful''. Planet MOA-2008-BLG-310-L b, a sub-Saturnian mass planet recently detected in the Galactic Bulge with the technique of microlensing (Janczak et al. 2009), certainly inspires this feeling of transcendence Poincar\\'e describes. But its name hardly helps on conveying it. One of the main reasons I consider for naming the extrasolar planets is the Copernican Principle itself. Our place in the cosmos is not special in any way, so there is no reason why only the planetary objects in the solar system should be named. Shakespeare would perhaps disagree with me and say that Io by any other name would smell as bad; and it is true that HD 128311 b will have the same radial velocity curve irrespectively of us naming it after a catalog number or after Bacchus. However, the non-special nature of our place in the Universe is better underscored by naming our neighbors. Mercury - Venus - Earth - Mars is a sequence of equals. Sol b - Sol c - Earth - Sol d would implicitly imply that the Earth is special in some way. Likewise, Jupiter is being paired to obscure names such as XO-1 b, TrES-4 b, and OGLE-TR-182 b, which does not help educators convey the message that these planets are quite similar to Jupiter. In stark contrast, the sentence``planet Apollo is a gas giant like Jupiter'' is heavily - yet invisibly - coated with Copernicanism. One reason given by the IAU for not considering naming the extrasolar planets is that it is a task deemed impractical. One source is quoted as having said ``if planets are found to occur very frequently in the Universe, a system of individual names for planets might well rapidly be found equally impracticable as it is for stars, as planet discoveries progress.'' {\\footnote{\\url{http://www.iau.org/public_press/themes/extrasolar_planets/}}}. This leads to a second argument. It is indeed impractical to name all stars. But some stars are named nonetheless. In fact, all other classes of astronomical bodies are named. Galaxies are named. Stars are named. Even asteroids are named. So why not name the exoplanets? Granted, not all galaxies have names. Only the nearby ones, and in this case, the names are easily fixed due to the shape of the galaxy - Whirlpool, Antenna, Sombrero -, or of the constellation - Andromeda, Circinus, Carina. Star naming also has a criterion - the brightness. All stars brighter than $V$=1.5 have a proper name, the frequency of named stars declining as the magnitude increases{\\footnote{S. Boscardin, from Observat\\'orio Nacional, Rio de Janeiro, Brazil, pointed to me that there are four unnamed stars brighter than $V$=2.5. They are gamma and delta Velorum, of magnitudes $V$=1.8 and 2.0, and epsilon and eta Centauri, both of magnitude $V$=2.5. I credit him here for this information, and echo his suggestion that the IAU should consider naming these four lone stars. He further informs me that there are 84 stars between 2.5$<$$V$$<$3.0, 51 of which are named; and 106 stars between 3.0$<$$V$$<$3.5, 33 of which are named.}}. Yet nothing has stopped people from naming over 15,000 asteroids and minor planets. In fact, it seems to be that the main reason for over 400 known exoplanets remaining unnamed is that no one has yet done the job of naming them all. Indeed, as discoveries proceed (and hopefully skyrocket with the Gaia mission), naming all planets will be impractical. But the benefits of having some of them named is as clear as in the case of stars. In this manuscript, I set myself to the task. In some cases, planets have already been nicknamed. For some, the epithet is sound. On naming 51 Pegasi b, I immediately thought of Bellerophon, the rider of the winged horse. Then I found out that someone else also shared the same taste for mythological associations and had already nicknamed the planet with the same name. I also came across the webpage of the ``Extrasolar planet naming society'', created in November 2008 with the idea of organizing a concerted collective effort to name the known exoplanets. Unfortunately, it did not seem to have gained any momentum. Individual efforts seem to beat collective ones in this case. I highlight here the webpage of Devon Moore, where over 40 exoplanet names have been suggested {\\footnote{\\url{http://nuclearvacuum.wikia.com/wiki/Names_for_extrasolar_planets}}}. Some of the names Moore chose are indeed good, some I have reasons to object. I will go back to that later. Discoverers have also attempted to name the fruit of their labor. The three planets around Upsilon Andromedae have been nicknamed ``Fourpiter'', ``Twopiter'', and ``Dinky''. These names were suggested by a class of 4th graders, and have the advantage of carrying information on the planets' size. However, they are tantalizingly unpalatable to those of a more classical mind. No offense to the very creative children, but apparently long gone are the times of Venitia Burney, then 11 years old, who suggested the name of Pluto, the Roman-Greek god of the underworld (not the Disney dog), for the distant, cold, (ex-)planet discovered in 1930. 70 Virginis b was similarly nicknamed ``Goldilocks'', for its position in the habitable zone of its star. One could imagine that had the discovery been made by a Swedish team if they would have called the planet {\\it Lagom}, after their unique word for ``just right''. Not to mention the internal names used by discovery teams that sometimes leak to the media, such as ``Xena'', ``Easterbunny'' and ``Santa'' (the Kuiper Belt objects Eris, Makemake, and Haumea, respectively). In this manuscript, I advocate a return to classic tradition, and propose a simple way to name the known exoplanets, exposed in Sect 2. It basically consists of giving them names from Roman-Greek mythology associated with the constellation that the host star pertains. The columns in Table~1 show, respectively, the names of the planets listed by constellation, the name used in the astronomical literature, their masses in units of Jupiter's mass ($M_J$), semi-major axes in astronomical units (AU), eccentricities, right ascension, declination, and, finally, the proposed name{\\footnote{Table~1 can also be found online at \\\\\\url{http://www.mpia.de/homes/lyra/planet_naming.html}}}. The mythological associations for these names are explained in Sect 3. The method used was the following. I got the planets from Jean Schneider's {\\it The Extrasolar Planets Encyclopaedia}, and listed the planets by constellation. I then suggest names for the planets with the help of a dictionary of Greek mythology (Dixon-Kennedy 1998) and extensive use of Wikipedia. It is commonplace that information in Wikipedia is not always trustable, and has to be used with care. That is correct, but when well-handled, the information there available is vital. For instance, listing the planets by constellation would have been quite time-consuming without Wikipedia, since the papers not always mention it, giving only the coordinates. To my surprise and amazement, almost all extrasolar planets have a wiki, that also informs in which constellation the star is located. I also made use of the following online source, \\url{http://www.theoi.com/}, that works as an online interactive dictionary of Greek mythology. ", "conclusions": "In this manuscript I suggest names for the 403 extrasolar planet candidates known as of October 2009. The suggestion is based on the classical tradition of giving names from Roman-Greek mythology to astronomical bodies. The association with the myths is in many cases purposively loose, to enable more flexibility on naming further planets as they are discovered. As said in \\sect{subsect:nonroman}, the system does not exclude other mythologies, which may be used if a suitable association with the constellation can be established. The system also has some power of prediction. We can say for instance, that the next planet discovered in Eridanus could be called Minos, after one of the three judges of Hades. Indeed, in a former version of this manuscript I had suggested that the next three planets could be called Minos, Eachus, and Radhamantus. Then, on Sep 17, the Encyclopedia of Extrasolar Planets was updated. One of the new planets was in Eridanus, and I promptly added it to Table 1 as Radhamantus. In the update of Oct 9, another planet in Eridanus was announced. I added it to Table 1 as Eachus. Along the course of writing this manuscript, I could already experience another small benefit of having the planets named. Scrolling the long table up and down, it was much easier to remember ``Typhon'' than ``HD 168443 c'' I further stress that the system proposed here does not intend to supplant the one in vogue. It is not a change, but an addition. Stars are known by many names. Merope is also known as 23 Tau, HD 23480, HIC 17608, HR 1156, 2MASS J03461958+2356541, and V971 Tau, to name a few. The current naming scheme of assigning minor letters to the names of stars will of course be kept for scientific publications, much in the same way that we use HD 135742 instead of Zubenelschamali and NGC 4755 instead of Jewel Box. The proper name is a bonus aimed at popular writings. One drawback I can think of is that it may lead the public to assume that the constellations are somehow physical. In some way, the misconception already exists. I recall this anecdote, for which I unfortunately cannot find the reference, that a piece of popular scientific writing once defined a constellation as ``a group of stars. Up to date, astronomers have found only 88''. On the other hand, we can invert the argument and see it as a golden opportunity to fight this misconception. Along with the names and their associations, it has to be pointed that the constellations are human invention and just a useful way of mapping the sky. Naming the planets after the myth of the constellations is no more misleading than using stellar names such as 14 Herculi, 70 Virginis or Upsilon Andromedae. It should also be pointed that this manuscript was sent to the IAU Commission 53 on exoplanets, whose majority still opposes the idea of adding names to planets. The strongest concern of the commission had to do with the definition of an exoplanet. The Encyclopedia of Extrasolar Planets, complete as it is, lists candidate planets, and some of them are not confirmed. Therefore, some of the candidates assigned names here may as well be just low luminosity stars, brown dwarfs, a stellar spot or, as noted by a member of the commission, even a mote of dust in the spectrograph. In that case, the name should be withdrawn and re-assigned to other, confirmed, planet. \\vspace{1cm} {\\it Acknowledgments.} I thank the reading and comments of J. Alves, S. Boscardin, A. Johansen, M.-M. Mac Low, A. Moitinho, N. Piskunov, H. Rocha-Pinto, S. Soter, and P. Tsalmantza, that helped me on identifying the main points of critique. Interestingly, Portuguese speakers expressed concern regarding the use of the Lusiad. Westerns were concerned with the supposed Eurocentrism of mainly using Roman-Greek myths. It is great to see that we are living in multicultural times and that instead of trying to inflate national prides and highlighting boundaries, we are actually trying to distance ourselves from them. I also thank M. Schmitz for organizing an unorthodox peer review among the IAU Comission 53 and playing the role of editor. \\vspace{1cm}" }, "0910/0910.0278_arXiv.txt": { "abstract": "Non-radial pulsations (NRPs) are a proposed mechanism for the formation of decretion disks around Be stars and are important tools to study the internal structure of stars. NGC 3766 has an unusually large fraction of transient Be stars, so it is an excellent location to study the formation mechanism of Be star disks. High resolution spectroscopy can reveal line profile variations from NRPs, allowing measurements of both the degree, $l$, and azimuthal order, $m$. However, spectroscopic studies require large amounts of time with large telescopes to achieve the necessary high S/N and time domain coverage. On the other hand, multi-color photometry can be performed more easily with small telescopes to measure $l$ only. Here, we present representative light curves of Be stars and non-emitting B stars in NGC 3766 from the CTIO 0.9m telescope in an effort to study NRPs in this cluster. ", "introduction": "Be stars are a class of non-supergiant B-type stars with Balmer and other line emission features due to an equatorial decretion disk. The disk is likely the result of a combination of the star's rapid rotation (near the critical limit) and non-radial pulsations (NRPs; Porter \\& Rivinius 2003). NRPs are spherical harmonic waves traversing the surface of a star. These pulsations can be found in multiple frequencies on the surface simultaneously (Rivinius, Baade, \\& \\u Stefl 2003). There are two primary classes of NRP modes: $g$- and $p$-modes. $g$-modes are described by a low frequency pulsation that has gravity as its restoring force. The dominant oscillation in this mode is transverse across the surface. $p$-modes are dominated by high frequency, radial oscillations with a pressure restoring force (De Ridder 2001). These modes in main-sequence, pulsating B stars are driven by the $\\kappa$ mechanism (Guti\\'errez-Soto et al. 2007). Temperature and flux gradients are established between the dimmer, cooler material on the peaks of the pulsations and the brighter, warmer material in the troughs. The flux variations over the stellar surface are then observed as either ripples within photospheric absorption line profiles or as periodic variations in magnitude. A large, high-resolution spectroscopic study would reveal both the degree, $l$, and the azimuthal order, $m$, but such studies are challenging due to the need for large amounts of time on a large telescope. Photometry, which is easily performed with data gathered by small telescopes, only measures $l$ (Rivinius, Baade, \\& \\u Stefl 2003). McSwain et al. (2008) previously showed that NGC 3766, an open cluster in Centaurus, is rich with transient Be stars. In an effort to detect NRPs and study the formation of these transient disks, we are currently performing a long-term photometric study of the cluster. Here we present preliminary differential light curves that reveal magnitude variations of several Be stars that are consistent with NRPs. ", "conclusions": "" }, "0910/0910.2757_arXiv.txt": { "abstract": "We have performed C$^{18}$O ($J$=1--0) mapping observations of a $20'\\times20'$ area of the OMC-1 region in the Orion A cloud. We identified 65 C$^{18}$O cores, which have mean radius, velocity width in FWHM, and LTE mass of 0.18$\\pm$0.03 pc, 0.40$\\pm$0.15 km s$^{-1}$, and 7.2$\\pm$4.5 $M_\\odot$, respectively. All the cores are most likely to be gravitationally bound by considering the uncertainty in the C$^{18}$O abundance. We derived a C$^{18}$O core mass function, which shows a power-law-like behavior above 5 $M_\\odot$. The best-fit power-law index of $-2.3\\pm0.3$ is consistent with those of the dense core mass functions and the stellar initial mass function (IMF) previously derived in the OMC-1 region. This agreement strongly suggests that the power-law form of the IMF has been already determined at the density of $\\sim10^{3}$ cm$^{-3}$, traced by the C$^{18}$O ($J$=1--0) line. Consequently, we propose that the origin of the IMF should be searched in tenuous cloud structures with densities of less than 10$^{3}$ cm$^{-3}$. ", "introduction": "\\label{introduction} One of the most important observational features of the stellar initial mass function (IMF) is its power-law-like nature above 1 $M_\\odot$, as $dN/dM \\propto M^{-\\gamma}$. In the solar neighborhood, the power-law index $\\gamma$ seems to be greater than 2 \\citep{sal55,kro01a}, which characterizes the statistical properties of stars. In particular, both the total number and mass of stars are dominated by those of low-mass stars of $\\sim$ 1 $M_\\odot$. It is natural to consider that the origin of the IMF shape is related to the mass distribution of its natal gas in molecular clouds. Many authors have investigated dense gas ($10^{4-5}$ cm$^{-3}$) in molecular clouds by using (sub)millimeter dust continuum emission and/or molecular line emissions having high critical densities such as the H$^{13}$CO$^{+}$($J$=1--0) and N$_{2}$H$^{+}$($J$=1--0) lines \\citep[e.g.,][]{mot98,rei06,nut07,ike07,wal07,eno08}. They identified numerous cores that have typical sizes of 0.05 -- 0.1 pc and masses of 1 -- 10 $M_{\\odot}$ in nearby ($\\leq$ 500 pc) star forming regions such as Orion, Ophiuchus, Perseus, and Serpens. The molecular line studies showed that the cores are gravitationally bound and are likely to produce stars. Moreover, they found that the core mass functions (CMFs) derived by using the dense gas tracers, referred to as DCMFs hereafter, seem to have power-law-like behaviors in high-mass parts, whose $\\gamma$ values are very similar to that of the IMF. One exception is that \\citet{kra98} found a significatly smaller $\\gamma$ value of 1.7 for the S140 and M17SW regions by using the C$^{18}$O($J$=2--1) line with a high critical density comparable to that of H$^{13}$CO$^{+}$(1--0). Considering that the power-law form of the IMF has been already determined at the formation stage of the cores with the densities of $10^{4-5}$ cm$^{-3}$, it is likely that more tenuous structures of molecular clouds have a key to understanding the origin of the power-law nature of the IMF. It has been suggested that the mass functions in the tenuous gas structures % are different from the DCMFs. \\citet{kra98} carried out a systematic study of the mass functions in the tenuous gas structures of 10$^{3}$ cm$^{-3}$ or less by using the $^{12}$CO(2--1), $^{13}$CO(1--0; 2--1), and C$^{18}$O(1--0) lines in various molecular clouds. They showed that their mass functions seem to have a common power-law form, and the $\\gamma$ value of 1.7$\\pm$0.1 is significantly smaller than those of the DCMFs and the IMF. \\citet{hei98} derived $^{12}$CO($J$=1--0, 2--1) mass functions in the MCLD 123.5$+$24.9 and Polaris Flare regions and found that $\\gamma$ = $1.8\\pm0.1$. \\citet{won08} found $\\gamma$ = 1.7 in the C$^{18}$O($J$=1--0) mass function of RCW 106. These facts mean that the power-law index of the IMF may be originated in the formation process of the dense gas of $10^{4-5}$ cm$^{-3}$ from the tenuous gas of $10^{2-3}$ cm$^{-3}$. However, one should be careful in comparing the $\\gamma$ value of the tenuous gas mass function to those of the DCMFs and the IMF. This is because the tenuous gas mass functions described above were derived by the spatial resolutions larger than 0.1 pc, which cannot resolve star-forming cores, and/or by using optically-thick tracers such as $^{12}$CO and $^{13}$CO. To fairly compare the tenuous gas mass function with the DCMFs and the IMFs, one should achieve higher spatial resolutions than 0.1 pc and use optically-thin tracers. In this paper, we present a CMF derived by C$^{18}$O($J$=1--0) mapping observations of the OMC-1 region, which is located at the center of the Orion A Giant Molecular Cloud. The aim of this study is to examine whether or not the common power-law form between the DCMFs and the IMF has been already determined in the tenuous gas of $\\leq10^{3}$ cm$^{-3}$ by focusing on the power-law index $\\gamma$ in the high-mass part of the CMF. The C$^{18}$O($J$=1--0) molecular line emission is suitable for deriving the CMF \\citep[e.g.,][]{tac02} because the line has a relatively small critical density of $\\sim$10$^{3}$ cm$^{-3}$ \\citep{ung97} and is typically optically thin (see \\S \\ref{coreidentification}). The OMC-1 region is one of the best regions to investigate the CMF, because the IMF of the associated Orion Nebula Cluster (ONC) has been derived \\citep{hil97,mue02}, and because the power-law form of the H$^{13}$CO$^{+}$ DCMF is shown to be quite similar to that of the ONC IMF by \\citet{ike07}. Furthermore, at the distance to the Orion A cloud of 480 pc \\citep{gen81} we can easily resolve the cores with radii of $\\sim$0.1 pc. As described in \\S \\ref{observation}, the mapping observations have been done with the effective spatial resolution of 26$''$.4 ($=$ 0.06 pc), which is high enough to resolve the dense cores in the OMC-1 region \\citep{ike07}. ", "conclusions": "" }, "0910/0910.0108_arXiv.txt": { "abstract": "The open clusters' fundamental physical parameters are important tools to understand the formation and evolution of the Galactic disk and as grounding tests for star formation and evolution models. However only a small fraction of the known open clusters in the Milky Way has precise determination of distance, reddening, age, metallicity, radial velocity and proper motion. One of the major problems in determining these parameters lies on the difficulty to separate cluster members from field stars and to assign membership. We propose a decontamination method by employing 2MASS data in the encircling region of the clusters NGC1981, NGC2516, NGC6494 and M11. We present a decontaminated CMD of these objects showing the membership probabilities and structural parameters as derived from King profile fitting. ", "introduction": " ", "conclusions": "" }, "0910/0910.2611_arXiv.txt": { "abstract": "{Using numerical ray tracing, the paper studies how the average distance modulus in an inhomogeneous universe differs from its homogeneous counterpart. The averaging is over all directions from a fixed observer not over all possible observers (cosmic), thus is more directly applicable to our observations. In contrast to previous studies, the averaging is exact, non-perturbative, and includes all non-linear effects. The inhomogeneous universes are represented by Swiss-cheese models containing random and simple cubic lattices of mass-compensated voids. The Earth observer is in the homogeneous cheese which has an Einstein - de Sitter metric. For the first time, the averaging is widened to include the supernovas inside the voids by assuming the probability for supernova emission from any comoving volume is proportional to the rest mass in it. Voids aligned along a certain direction give rise to a distance modulus correction which increases with redshift and is caused by cumulative gravitational lensing. That correction is present even for small voids and depends on their density contrast, not on their radius. Averaging over all directions destroys the cumulative lensing correction even in a non-randomized simple cubic lattice of voids. At low redshifts, the average distance modulus correction does not vanish due to the peculiar velocities, despite the photon flux conservation argument. A formula for the maximal possible average correction as a function of redshift is derived and shown to be in excellent agreement with the numerical results. The formula applies to voids of any size that: (1) have approximately constant densities in their interior and walls; and (2) are not in a deep nonlinear regime. The average correction calculated in random and simple cubic void lattices is severely damped below the predicted maximal one after a single void diameter. That is traced to cancellations between the corrections from the fronts and backs of different voids. The results obtained allow one to readily predict the redshift above which the direction-averaged fluctuation in the Hubble diagram falls below a required precision and suggest a method to extract the background Hubble constant from low redshift data without the need to correct for peculiar velocities.} ", "introduction": "It is fair to say that the standard cosmological $\\Lambda$CDM model is facing a phenomenological crisis. The dark matter carrier has been evading direct detection for decades and the origin of dark energy remains a theoretical puzzle. The most natural candidate is the vacuum state energy, but the flat-space Quantum Field Theory is incapable of calculating it, producing an estimate that is $10^{120}$ times as large as the value suggested by the supernova observations. Presumably, that would be rectified in a fully quantized theory of gravitation which unfortunately does not yet exist. Remaining within the standard General Relativity, there are attempts to explain dark energy as an apparent quantity arising from the averaging procedure that maps the real lumpy universe to the idealized homogeneous Friedman-Lemaitre-Robertson-Walker (FLRW) metric. Such approaches separate into three distinct classes. First, one could average the Einstein equations over spatial hypersurfaces and hope the analogue of the Friedman equation contains a significant effective $\\Lambda$ term. Such an idea is conceptually problematic since inhomogeneous universes do not naturally pick up preferred spatial slices to average over, unlike the homogeneous FLRW models. It is shown in \\cite{wald} that the result of such averaging depends on the arbitrary choice of slicing and has no coordinate independent meaning - even Minkowsky spacetime could produce an apparent acceleration. A cousin to the first approach is averaging over the null hypersurface of the past light cone of the observer, \\cite{lightcone, valerioth}. It is not proven that such averages have a physical meaning and are not simply coordinate quantities. In the second approach, the second order perturbations to the Einstein equations are calculated and treated as an effective energy momentum tensor term. Unfortunately, as pointed out in \\cite{wald}, the second order perturbation term is not gauge independent and therefore is not suitable to play the role of an energy momentum tensor in the Einstein equation. The present paper belongs to the third approach which is completely coordinate independent since it calculates relations only between physically observable quantities such as redshift and luminosity distance (or its logarithmic measure - the distance modulus). Papers of this type, for example \\cite{norwegian}, demonstrate that the distance modulus - redshift Hubble diagram of type Ia supernovas can be reproduced without any dark energy for an observer occupying the center of a giant ($\\sim$ Gpc) underdense spherical bubble. That is possible because the central observer measures a local Hubble parameter that is larger than the global one. The corresponding shift on the Hubble diagram between the true background and the one assumed by the observer allows for fitting the supernova data. Bubbles of such a large scale have significant peculiar velocities away from their center. That creates a fine tuning problem - the observer has to be in the vicinity of the bubble center \\cite{offcenter} to observe a small CMB dipole equal to the measured $v/c\\sim 0.002$ (Local Group velocity $\\approx620$ km/s). This violates the Copernicus principle that we do not occupy a special position in the universe. Other papers of the same class studied the collective effect of a configuration of many bubbles/voids of smaller size, thus avoiding big peculiar velocities and significant anisotropies in CMB. That scenario can be conveniently simulated in the \"Swiss-cheese\" toy model \\cite{lemaitre, plebanski, bondi} which has the virtue of being analytically solvable. It is constructed by removing spherical regions from a homogeneous FLRW background (the \"cheese\") and replacing them with inhomogeneous density distributions with the same gravitating mass (mass-compensated voids). It was used in \\cite{valerio, valerioth} for an observer looking along the diameters of a string of aligned voids of radius $350$ Mpc. A positive cumulative change in the distance modulus was found which increased with redshift with respect to the homogeneous background. Although the size of the correction was significant at high redshift, it was not large enough to fit the supernova Hubble diagram and it substituted only partially for dark energy. A smaller correction due to aligned voids with a more mundane radius of $23 h^{-1}$ Mpc was calculated in \\cite{brouzakis-eqs}. The result obtained in \\cite{valerio} was criticized in \\cite{vanderveld} where it was demonstrated to stem from the cumulative weak lensing defocusing of the rays passing diametrically through the aligned voids. The same paper showed that the correction to the distance modulus vanishes when averaged over randomized impact parameters of the rays entering the voids. This conclusion is in accord with an old argument \\cite{weinberg} based on the gravitational lensing conservation of the total photon flux. The implicit geometrical assumptions in \\cite{weinberg} have been challenged in more recent papers \\cite{mustapha, rose} (and references therein) and the present work will demonstrate they are violated close to the observer due to peculiar velocity modifications of the redshift surfaces. The studies mentioned above are deficient in several areas. Averaging over directions was not performed in \\cite{valerio}. The calculations in \\cite{vanderveld} were done using weak lensing theory neglecting the time dependence of the void density and possible nonlinear effects which become significant at low redshifts. The averaging assumed that the ray impact parameters had a uniform probability distribution over the cross-section of a void. That is increasingly true at high redshifts but does not hold in the local neighborhood where the rays have to converge on the observer. More importantly, \\cite{vanderveld} averaged only over supernovas residing in the homogeneous cheese. The same was done in a study that obtained an analytical estimate of the corrections to the luminosity distance, \\cite{brouzakis}. Such a bias is unjustified, first because supernovas in the voids produce much bigger corrections than the ones in the cheese \\cite{valerio}, and second, the observed supernovas occur more frequently in denser regions like the void walls than in the cheese. The goal of the present paper is to address once again the question whether the direction-averaged distance modulus correction from a configuration of voids really vanishes but this time in the context of a calculation that: (1) is exact, non-perturbative, an includes all possible non-linear effects; (2) is armed with a physically sensible averaging procedure free of ad hoc assumptions about impact parameters \\cite{vanderveld} or random cancellations of corrections \\cite{weinberg}; and (3) does not neglect the supernovas inside voids. The final outcome of the calculations cannot be guessed on the grounds of the previous studies because it may depend on non-linear effects, the choice of averaging procedure, and the supernova selection, none of which was taken into account before. The calculations are performed in Swiss-cheese models with mass-compensated voids of two radii having observational support: $30$ and $300$ Mpc. The cheese is spatially-flat matter-only Einstein - de Sitter (EdS). The observer is placed in the cheese since at present there is no indication that we occupy a void. The first Swiss-cheese model is set up in section 2. The observer shoots past-directed light rays in all directions. The light propagation geodesic equations and the luminosity distance tracing along each ray are discussed in section 3. The physically sensible averaging procedure in this paper assumes that the probability for supernova emission from a comoving volume is proportional to the rest mass in it. Averaging the distance modulus for a single void is discussed in section 4 and it carries the essential characteristics of the procedure for many voids. The main outcome of that section is a simple formula to estimate the maximal average corrections to the distance modulus. Section 5 studies the cumulative correction due to gravitational lensing along a string of aligned voids. Numerical results from averaging within random and simple cubic void lattices are presented in section 6. The effect on the distance modulus produced by large voids of radius $300$ Mpc, which leave a measurable imprint on CMB, is evaluated in Section 7. The summary and conclusions section discusses the answer to the main question addressed by the present study and the practical importance of the obtained low-redshift results for future surveys. The speed of light in this paper is $c=1$, and times and distances are measured in megaparsecs (Mpc), occasionally giving time in megayears (Myr) for convenience. The gravitational constant $G$ is kept explicit in all equations so that the reader can easily substitute with a favorite value. The usual geometrized units, $G=1$, were used in the numerical calculations. ", "conclusions": "The paper studied how the distance modulus, averaged over all lines of sight in an inhomogeneous universe, differs from its homogeneous counterpart. The inhomogeneities were represented by random and regular lattices of mass-compensated voids in Swiss-cheese models. Two void layouts were considered with a small ($30$ Mpc) and a big ($300$ Mpc) radius, the first observed in the SDSS data \\cite{visualvoid} and the second deduced from its imprint on CMB \\cite{granett}. The Earth observer was put in the cheese but the conclusions will not change significantly for an observer inside one of the voids. For the first time, the averaging of the distance modulus was widened to include the supernovas inside voids. That was made possible by assuming that the probability of supernova emission from a comoving volume is proportional to the rest mass in it. Unlike previous studies \\cite{weinberg, vanderveld}, the average correction to the distance modulus was calculated by exact numerical ray tracing without using assumptions about the ray impact parameters or perturbation theory approximations, and the result includes all linear and non-linear effects. The distance modulus correction, due to gravitational lensing, that accumulates along a ray crossing diametrically a sequence of aligned voids \\cite{valerio, brouzakis-eqs, vanderveld} was calculated. The correction increases linearly and becomes quite significant at high redshifts, see Fig.\\ref{cumucorr}. It is roughly proportional to the density contrast in the void interior but does not depend on the void radius. It was demonstrated that the distance modulus, averaged over all directions in a lattice of voids, does not show a cumulative correction at high redshifts. That is true not only in a random lattice (see Fig.\\ref{randvoidsave}) but also in a regular simple cubic one (see Fig.\\ref{scvoidsave} and Fig.\\ref{scvoidsave300}) implying that void randomization is not necessary for the destruction of the cumulative correction when averaging, contrary to popular belief. A large cumulative correction is still observed along special directions in a regular void lattice \\cite{valkenburg}, but the probability for the required void alignment is vanishing in the real universe. At low redshifts $z$, perturbation theory predicts that the line-of-sight peculiar velocities $v_r$ are capable of producing significant fluctuations in the luminosity distance $\\Delta d_L(z)/d_L \\approx - v_r/z$ \\cite{hui, haugbolle} ($v_r$ measured in speed of light units). That leads to a scatter in the local Hubble diagram $\\Delta H/ H = -\\Delta d_L/d_L$ and, by (\\ref{fraccorr}), to a non-vanishing average correction $\\left<\\Delta \\mu(z)\\right>\\; \\approx \\, -2.17 \\left<\\Delta H / H \\right> \\, \\approx \\, - 2.17 \\left/\\,\\, z \\, $, where the so called peculiar velocity monopole $\\left$ is the line-of-sight peculiar velocity at redshift $z$, averaged over all directions. A measure of the fluctuation magnitude is its average within a sphere of radius $R$ around the observer, $[\\Delta H/H]_R$, which varies with location. Its cosmic variance over all possible observers/locations has been calculated previously in several cosmological models by N-body simulations \\cite{edwin} and by linear perturbation theory using the matter power spectrum \\cite{shiturner, shidursi}. Such variances are employed to predict confidence intervals for the value of $[\\Delta H/H]_R$ measured by a random observer. Since the obtained cosmic variance of $[\\Delta H/H]_R$ decreases with $R$ \\cite{shiturner, shidursi}, with the cosmic mean being zero, the likelihood to measure large average fluctuations decreases with redshift, as expected. % The cosmic variance of the sphere-average $[\\Delta H/H]_R$ over all observers cannot be used to estimate the average over all directions $\\left<\\Delta H(z)/H\\right>$ in the redshift bin $z$, experienced by a \\textit{particular} observer. The results in the present paper allow to calculate $\\left<\\Delta H(z)/H\\right>$ by studying such an observer in a simple model of our local neighborhood. The averages and variances so obtained are not cosmic, thus are more applicable to our astronomical observations. They cannot be calculated from a given matter power spectrum but require a concrete numerical realization of an inhomogeneous universe. To the best knowledge of the author, this is the first time when the non-vanishing direction-averaged correction $\\left<\\Delta \\mu(z)\\right>$ is calculated as a function of redshift for universes containing voids. The upper bound (\\ref{corr}) for $\\left<\\Delta \\mu(z)\\right>$ was motivated using linear perturbation theory and was found to agree with the numerical calculations on Fig.\\ref{maxcorr} and Fig.\\ref{maxcorr300}. The parameter $\\eta$ depends on the particular form of inhomogeneities and the choice of averaging procedure. The numerically obtained value $\\eta \\sim 0.20$ indicates that the maximal average correction is $20 \\%$ of the naive correction corresponding to the maximal peculiar velocity. This $\\eta$ turned out to be a universal constant, approximately independent of the void size, for voids which have approximately constant densities in their interior and walls that are not in a deep nonlinear regime. Due to void randomization, it is expected that the average correction will be suppressed below the maximal $20 \\%$ and will tend to zero. The efficiency of that process and at what redshift it sets in has not been studied before. The calculations revealed that the average correction within void lattices is indeed severely damped below the predicted maximal average correction due to cancelations between the fronts and backs of different voids. As figures \\ref{randvoidsave}, \\ref{scvoidsave}, and \\ref{scvoidsave300} demonstrate, the cancelation is surprisingly efficient at low redshifts even in regular lattices and the average correction drops below $0.01$ mag after a single void diameter. Nevertheless, the average correction is not zero close to the observer, indicating that the implicit assumptions of the photon flux conservation argument stated in \\cite{weinberg} do not apply at low redshifts. The nonzero $\\left<\\Delta \\mu(z)\\right>$ is observable, provided there are enough supernovas $N_z$ in redshift bin $z$ to reduce the sampling statistical error of the average: $\\sigma[\\,\\left<\\Delta \\mu(z)\\right>\\,] = \\sigma[\\Delta\\mu(z)] /\\sqrt{N_z}$, where $\\sigma[\\Delta\\mu(z)]$ is shown as a thin solid curve on figures \\ref{randvoidsave}, \\ref{scvoidsave}, and \\ref{scvoidsave300}. An underdense \"Hubble bubble\" around us with a Hubble parameter slightly higher than the global one (correspondingly $\\left<\\Delta \\mu(z)\\right> < 0$) has been argued for in \\cite{zehavi, jha}. However, that claim was not confirmed in \\cite{giovanelli, neill} and was shown to depend on the way the supernova magnitudes were extracted from the photometric data \\cite{hicken}. Local bubble or not, fluctuations in the distance modulus binned average begin to appear in the newest data sets, Fig.20 in \\cite{kessler}, and Fig.7 in \\cite{sandage}. Significant efforts in contemporary cosmology are aimed at measuring a possible time evolution in the dark energy equation of state. That would require fixing the Hubble constant at low redshifts to a $1 \\%$ precision \\cite{jha, riess}. Ongoing and future low redshift surveys will boost the number of the observed nearby supernovas sufficiently to bring the sampling error of the average $\\sigma[\\left<\\Delta H / H \\right>]\\approx \\; \\sigma[\\,\\left<\\Delta \\mu(z)\\right>\\,] \\, / \\, 2.17 $ below $1 \\%$. In contrast, the directional average itself $\\left<\\Delta H / H \\right> \\approx \\;- \\left<\\Delta \\mu(z)\\right> \\, / \\, 2.17 $, generated by the coherent peculiar motion, is not influenced by the supernova number and will degrade significantly the errors in measuring the dark energy time dependence \\cite{hui, cooray}. Consequently, the peculiar velocities need to be subtracted from the low redshift Hubble diagram. Several methods are utilized for that. The Local Group (LG) of galaxies moves as a whole with respect to CMB at a speed $v_{LG}$, the so called peculiar velocity dipole \\cite{kocevski, watkins}. The most common way to correct the Hubble diagram for that motion is to adjust the observed redshifts to an observer riding the LG barycenter. The luminosity distance fluctuations seen in that frame are $\\Delta d_L/d_L \\approx -(v_r-v_{LG})/z$ \\cite{hui} and the velocity dipole will cancel out since the coherent velocity component of a nearby galaxy with respect to CMB is $v_r^{coh}\\approx v_{LG}$. Not correcting the CMB redshifts to the LG frame is permissible for a large supernova sample that covers densely and uniformly all directions: the coherent dipole fluctuation seen in the CMB frame is $\\Delta d_L/d_L \\approx- v_r/z\\approx -v_{LG}\\; cos(\\theta)\\, /z$ and that expression vanishes when averaged over the full solid angle. A further refinement is to correct the LG redshifts for infall velocities in the LG frame caused by the nearby superclusters represented by a crude linear multi-attractor model. That was done in the Hubble Key Project \\cite{mould} which measured the Hubble constant to a $9\\, \\%$ precision, but it was not very efficient at reducing the scatter in the low redshift Hubble diagram as seen on Fig.4 of \\cite{finalhubble}. Accordingly, the philosophy of the Hubble Key Project was to extract the Hubble constant mainly from secondary distance indicators at high redshifts. A finer method to correct for peculiar velocities is to calculate them from the observed galaxy distribution utilizing linear perturbation theory and choosing a galaxy-dark matter biasing parameter that minimizes the scatter on the Hubble diagram. That technique does reduce the scatter \\cite{radburn} but using it to infer the Hubble parameter from low redshift data produces a controversially high value of $H_0= 85 \\,km\\, s^{-1}\\,Mpc^{-1}$ \\cite{willick}. To avoid unreliable velocity corrections, many recent surveys \\cite{kessler, riess, sandagehubble} simply chose to include only objects above a certain redshift, $z>z_{min}$, where the peculiar velocities are considered insignificant compared to the Hubble flow. The Hubble constant and the dark energy equation of state parameter $w$ inferred from the low redshift data are sensitive to the choice of $z_{min}$ \\cite{jha, kessler}. As discussed in \\cite{kessler}, there is a wide variation in the values selected for $z_{min}$ in the literature, ranging from $0.01$ \\cite{sandagehubble} up to $0.023$ \\cite{riess}, which indicates that there is not a universally accepted prescription. Figures \\ref{randvoidsave}, \\ref{scvoidsave}, and \\ref{scvoidsave300} of the present paper allow to select $z_{min}$ readily. The required $\\Delta H/H = 0.01$ precision corresponds to $\\Delta \\mu_{goal} = 0.0217$. The sampling error of the average is $\\sigma[\\Delta\\mu(z)] /\\sqrt{N_z}$, where $\\sigma[\\Delta\\mu(z)]$ is given by the thin solid curves on the plots and $N_z$ is the number of objects/supernovas in redshift bin $z$. Provided there is sufficient statistics, the sampling error will fall below $\\Delta \\mu_{goal}$. In that case the limiting value $z_{min}$ is determined by the average $\\left<\\Delta \\mu(z)\\right>\\ $ itself - a natural choice is the redshift at which the maximal average correction (\\ref{corr}) (the dashed curves) drops below $\\Delta \\mu_{goal}$ or the redshift corresponding to a void diameter (where $\\left<\\Delta \\mu(z)\\right>\\ \\sim 0.01 $ due to front-back void cancellation), whichever is lower. In comparison, estimates based on the sphere-average $[\\Delta H/H]_R$ are more pessimistic: its cosmic variance in the $\\Lambda$CDM model drops below $1\\%$ at $z_{min}=0.05$ \\cite{shidursi} (linear estimate). If our neighborhood contains voids of the size observed on the CMB imprint \\cite{granett}, the required $z_{min}$ would be much larger, as Fig. \\ref{scvoidsave300} indicates. Excluding the objects below $z_{min}$ reduces the size of the sample and proportionally increases the statistical error \\cite{cooray}. The present paper suggests it is possible to preserve the low redshift data, instead of throwing it away, by collapsing it into an interval average $[\\Delta \\mu]$ over a redshift interval containing a void diameter. These averages turn out very close to zero: $[\\Delta \\mu]=0.003$ ($z=0\\, \\ldots \\,0.025$ on Fig.\\ref{randvoidsave}), $[\\Delta \\mu]=0.007$ ($z=0\\, \\ldots \\, 0.025$ on Fig.\\ref{scvoidsave}), $[\\Delta \\mu]=0.002$ ($z=0\\, \\ldots \\,0.25$ on Fig.\\ref{scvoidsave300}). That stems from the cancellation between positive and negative lobes in $\\left<\\Delta \\mu(z)\\right>$ and the mass averaging giving higher weights to higher redshifts where $\\left<\\Delta \\mu(z)\\right>$ is closer to zero. A vanishing $[\\Delta \\mu] \\approx 0$ means $[\\mu]\\approx[\\mu]^{EdS}$: \\begin{equation} \\label{extracthubble} \\sum_i W_i \\left<\\mu(z_i)\\right> \\approx \\sum_i W_i \\; \\mu_i^{EdS}(H_0, z_i), \\end{equation} where $W_i$ is the mass weight assigned to redshift bin $z_i$, $\\left<\\mu(z_i)\\right>$ is the distance modulus in that bin averaged over all directions, and the sum is over all the bins in the redshift interval. The lefthand side of the formula is calculated from the observational data. The weight $W_i$ is proportional to the rest mass inside the redshift bin which can be estimated as the corresponding mass in the homogeneous background model. Unlike a traditional Hubble diagram that weighs all redshift bins equally, here $W_i \\propto z_i^2 \\, \\Delta z_i$ at low redshifts. The righthand side of (\\ref{extracthubble}) depends solely on the Hubble parameter $H_0$ at low redshifts and on additional cosmological parameters of the background model at higher redshifts. It allows one to extract the value of $H_0$ from the low redshift data. This method will become possible for future surveys that: (1) have a dense full-sky coverage to calculate the average correction over all directions $\\left<\\Delta \\mu(z)\\right>$; and (2) contain a large number of objects in each redshift bin so that $\\left<\\Delta \\mu(z)\\right>$ is readily observable, not swamped by the sampling error. If the observer is inside a void, the average correction $\\left<\\Delta \\mu(z)\\right>$ starts with a negative lobe unlike the so-far considered case of an outside observer. A simple illustration of that is an observer at the center of a void, which will measure $\\Delta \\mu(z) = \\mu(z) - \\mu^{EdS}(z)<0$ for supernovas inside the void since the matter there is expanding faster than the background and it takes a smaller distance $d_L$ to achieve the same redshift. The interval average $[\\Delta \\mu]$ over a void diameter would still be very small since the mass averaging scheme adopted in this paper ascribes lower weights to low redshifts. Additional studies are needed to investigate the behavior of $[\\Delta \\mu]$ in the presence of superclusters - the other type of major inhomogeneities encountered in the universe. A plausible guess is that $[\\Delta \\mu]$ vanishes on a redshift interval encompassing a supercluster, justifying the validity of (\\ref{extracthubble}) in that case as well." }, "0910/0910.1612_arXiv.txt": { "abstract": "{We study the morphological content of a large sample of high-redshift clusters to determine its dependence on cluster mass and redshift. Quantitative morphologies are based on PSF-convolved, 2D bulge+disk decompositions of cluster and field galaxies on deep Very Large Telescope FORS2 images of eighteen, optically-selected galaxy clusters at $0.45 < z < 0.80$ observed as part of the ESO Distant Cluster Survey (``EDisCS''). Morphological content is characterized by the early-type galaxy fraction $f_{et}$, and early-type galaxies are objectively selected based on their bulge fraction and image smoothness. This quantitative selection is equivalent to selecting galaxies visually classified as E {\\it or} S0. Changes in early-type fractions as a function of cluster velocity dispersion, redshift and star-formation activity are studied. A set of 158 clusters extracted from the Sloan Digital Sky Survey is analyzed exactly as the distant EDisCS sample to provide a robust local comparison. We also compare our results to a set of clusters from the Millennium Simulation. Our main results are: (1) The early-type fractions of the SDSS and EDisCS clusters exhibit no clear trend as a function of cluster velocity dispersion. (2) Mid-$z$ EDisCS clusters around $\\sigma$ = 500 km/s have $f_{et} \\simeq$ 0.5 whereas high-$z$ EDisCS clusters have $f_{et} \\simeq$ 0.4. This represents a $\\sim$25$\\%$ increase over a time interval of 2 Gyrs. (3) There is a marked difference in the morphological content of EDisCS and SDSS clusters. None of the EDisCS clusters have early-type galaxy fractions greater than 0.6 whereas half of the SDSS clusters lie above this value. This difference is seen in clusters of all velocity dispersions. (4) There is a strong and clear correlation between morphology and star formation activity in SDSS and EDisCS clusters in the sense that decreasing fractions of [OII] emitters are tracked by increasing early-type fractions. This correlation holds independent of cluster velocity dispersion and redshift even though the fraction of [OII] emitters decreases from $z \\sim0.8$ to $z \\sim 0.06$ in all environments. Our results pose an interesting challenge to structural transformation and star formation quenching processes that strongly depend on the global cluster environment (e.g., a dense ICM) and suggest that cluster membership may be of lesser importance than other variables in determining galaxy properties. ", "introduction": "Our current paradigm for the origin of galaxy morphologies rests upon hierarchical mass assembly \\citep[e.g.,][]{steinmetz02}, and many transformational processes are at work throughout the evolutionary histories of galaxies. Some determine the main structural traits (e.g., disk versus spheroid) while others only influence properties such as color and star-formation rates. Disk galaxy collisions lead to the formation of elliptical galaxies \\citep{spitzer51,toomre72,farouki82,negroponte83,barnes92,barnes96,mihos96}, and the extreme example of this process is the build-up of the most massive galaxies in the Universe at the cores of galaxy clusters through the accretion of cluster members. Disks can also be transformed into spheroidals by tidal shocks as they are harassed by the cluster gravitational potential \\citep{farouki81,moore96,moore98}. Harassment inflicts more damage to low luminosity galaxies because of their slowly rising rotation curves and their low density cores. Galaxies can be stripped of their internal gas and external supply through ram pressure exerted by the intracluster medium \\citep{gunn72,larson80,quilis00}, and the result is a ``quenching'' (or ``strangulation'') of their star formation that leads to a rapid reddening of their colours \\citep[also see][]{martig09}. The task of isolating observationally the effects of a given process has remained a major challenge to this day. Many processes affecting galaxy morphologies are clearly environmentally-driven, and galaxy clusters are therefore ideal laboratories in which to study all of them. The dynamical state of a cluster, which can be observationally characterized by measuring mass and substructures, should be related to its morphological content. For example, the number of interactions/collisions suffered by a given galaxy should depend on local number density and the time it has spent within the cluster. Dynamically young clusters with a high degree of subclustering should contain large numbers of galaxies that are infalling for the first time. More massive clusters will contain more galaxies, but they will also have higher galaxy-galaxy relative velocities that may impede merging \\citep{lubin02}. Spheroidal/elliptical galaxies will preferentially be formed in environments where the balance between number density and velocity dispersions is optimal, but it is still not clear where this optimal balance lies. Cluster masses can be estimated from their galaxy internal velocity dispersion \\citep{rood72,dressler84,carlberg97,tran99,borgani99,lubin02}, through weak-lensing shear \\citep{kaiser93,schneider95,hoekstra00,clowe06} or through analysis of their hot X-ray emitting atmospheres \\citep[e.g., ][]{allen98}, and it will be used here as the main independent variable against which morphological content will be studied. The morphological content of high-redshift clusters is most often characterized by the fraction $f_{E+S0}$ of early-type galaxies they contain \\citep{dressler97,dokkum00, fasano00, dokkum01, lubin02, holden04, smith05, postman05, desai07,poggianti09b}. The bulk of the data available so far is based on visual classification. ``Early-type'' galaxies are defined in terms of visual classifications as galaxies with E or S0 Hubble types. A compilation of early-type fractions taken from the literature \\citep{dokkum00} shows a dramatic increase of the early-type fractions as a function of decreasing redshift from values around 0.4$-$0.5 at $z \\sim 1$ to values around 0.8 in the local Universe. However, the interpretation of this trend is not entirely clear as others \\citep[e.g.,][]{dressler97,fasano00,desai07,poggianti09b} have reported that the fraction of E's remains unchanged as a function of redshift and that the observed changes in early-type fractions are entirely due to the S0 cluster populations. S0 populations were observed to grow at the expense of the spiral population \\citep{smith05,postman05,moran07,poggianti09b} although others \\citep[e.g.,][]{holden09} have argued for no evolution in the relative fraction of ellipticals and S0s with redshift. \\citet{smith05} and \\citet{postman05} show that the evolution of $f_{E+S0}$ is in fact a function of both lookback time (redshift) and projected galaxy density. They find $f_{E+S0}$ stays constant at 0.4 over the range 1 $< t_{lookback} <$ 8 Gyr for projected galaxy densities $\\Sigma <$ 10 Mpc$^{-2}$. For high density environments ($\\Sigma$ = 1000 Mpc$^{-2}$), $f_{E+S0}$ decreases from 0.9 to 0.7. At fixed lookback time, $f_{E+S0}$ varies by a factor of 1.8 from low to high densities at $t_{lookback}$ = 8 Gyr and by a factor of 2.3 at $t_{lookback}$ = 1 Gyr. The difference between low and high density environments thus increases with decreasing lookback time. Both studies indicate that the transition between low and high densities occurs at $0.6R_{200}$ ($R_{200}$ is the projected radius delimiting a sphere with interior mean density 200 times the critical density at the cluster redshift, see Equation~\\ref{radius200}). \\citet{postman05} also find that $f_{E+S0}$ does not change with cluster velocity dispersion for massive clusters ($\\sigma$ $>$ 800 km/s). The data for one of their clusters also suggest that $f_{E+S0}$ decreases for lower mass systems. This trend would be consistent with observations of $f_{E+S0}$ in groups that show a strong trend of decreasing $f_{E+S0}$ versus decreasing $\\sigma$ \\citep{zabludoff98}. Finally, $f_{E+S0}$ seems to correlate with cluster X-ray luminosity at the 2-3$\\sigma$ level \\citep{postman05}. Recent works on stellar mass-selected cluster galaxy samples \\citep{holden07,vanderwel07} paint a different picture. The fractions of E+S0 galaxies in clusters, groups and the field do not appear to have changed significantly from $z \\sim 0.8$ to $z \\sim 0.03$ for galaxies with masses greater than 4$\\times 10^{10} M_{\\odot}$. The mass-selected early-type fraction remains around 90\\% in dense environments ($\\Sigma >$ 500 gal Mpc$^{-2}$) and 45\\% in groups and the field. These results show that the morphology-density relation of galaxies more massive than 0.5M$_{*}$ has changed little since $z \\sim 0.8$ and that the trend in morphological evolution seen in luminosity-selected samples must be due to lower mass galaxies. This is in agreement with \\citet{delucia04,delucia07} and \\citet{rudnick09} who have shown the importance of lower mass (i.e., fainter) galaxies to the evolution of the color-magnitude relation and of the luminosity function versus redshift. Another interesting result has come from attempts to disentangle age, morphology and environment in the Abell 901/902 supercluster \\citep{wolf07,lane07}. Local environment appears to be more important to galaxy morphology than global cluster properties, and while the expected morphology-density and age-morphology relations have been observed, there is no evidence for a morphology-density relation at a fixed age. The time since infall within the cluster environment and not density might thus be the more fundamental parameter dictating the morphology of cluster galaxies. A number of efforts have been made on the theoretical side to model the morphological content of clusters. \\citet{diaferio01} used a model in which the morphologies of cluster galaxies are solely determined by their merger histories. A merger between two similar mass galaxies produces a bulge, and a new disk may form through the subsequent cooling of gas. Bulge-dominated galaxies are in fact formed by mergers in smaller groups that are later accreted by clusters. Based on their", "conclusions": "\\label{discussion} In order to fully understand possible evolutionary trends observed here, it is important to determine how cluster velocity dispersion changes with redshift as a result of the hierarchical growth of structures. Are we looking at similar clusters when we focus on the same range of velocity dispersions in the SDSS and EDisCS clusters? \\citet{poggianti06} looked at the mean change in $\\sigma$ between $z$ = 0 and $z$ = 0.76 using a sample of 90 haloes from the Millennium Simulation uniformly distributed in log(mass) between 5 $\\times$ 10$^{12}$ and 5 $\\times$ 10$^{15}$ M$_{\\odot}$. Their Figure 8 shows how $\\sigma$ evolves over that redshift interval. For example, a $z$ = 0 cluster with $\\sigma$ = 900 km/s would typically have $\\sigma \\sim$ 750 km/s at $z$ = 0.76. This evolution is not sufficient to introduce biases in our analysis here. Indeed, selecting clusters with $\\sigma \\geq$ 600 km/s, say, at either $z$ = 0 or $z$ = 0.76 would keep nearly all the same clusters. Measured velocity dispersions may exhibit a large scatter with respect to the true halo mass particularly for low-mass clusters. The velocity dispersions for the SDSS and EDisCS clusters were calculated in a very similar way in order to minimize any biases. Velocity dispersions calculated from a small number of cluster members may be overestimates of the true cluster mass. Table~\\ref{clsample} lists 1103.7-1245b as the cluster with the lowest number of members ($N$ = 11). In order to check the robustness of our results, we re-ran our analyses by excluding SDSS clusters in Table~\\ref{sdss-cls-list} with $N < 10$ for which velocity dispersions may be less reliable and found that our results remained unchanged. \\citet{poggianti06} proposed a scenario in which two channels are responsible for the production of passive galaxies in clusters, and others \\citep{faber07,brown07} have proposed a similar scenario for the migration of galaxies from the \"blue cloud\" to the red sequence. \"Primordial passive galaxies\" are composed of galaxies whose stars all formed at very high redshift ($z >$ 2) over a short timescale. These galaxies have been observed in clusters up and beyond $z = 1$, and they largely comprise luminous ellipticals. \"Quenched passive galaxies\" have had a more extended period of star formation activity, and their star formation has been quenched after their infall into dense cluster environments. These quenched passive galaxies would then suffer the effects of cluster processes such as ram pressure stripping, harassment, strangulation and mergers to become S0 and earlier type galaxies. A key point of this scenario is that processes affecting morphology and star formation activity operate on different timescales as shown recently for the EDisCS sample by \\citet{sanchez09}. There is good evidence that star formation is quenched in galaxies over timescales of 1-3 Gyr after they have entered the cluster environment \\citep{poggianti99,poggianti06} whereas morphological transformation through mergers and harassment can take longer \\citep[$\\sim$ 5 Gyr,][]{moore98}. The best example of this is the fact that the vast majority of post-starburst galaxies in distant clusters, those that have had their star formation activity terminated during the last Gyr, still retain a spiral morphology \\citep{poggianti99}. Such a two-channel scenario would naturally explain observations indicating that the elliptical galaxy fraction actually remains constant with redshift while the S0 fraction rises with decreasing redshift \\citep{dressler97,fasano00,desai07}. Unfortunately, the VLT/FORS2 images do not have sufficient spatial resolution to disentangle E and S0 galaxies as mentioned in Section~\\ref{efrac-defn} to determine the exact contribution from each channel. We can therefore only study the overall production of early-type galaxies, but it should exhibit different behaviors with cluster global properties depending on the process(es) dominating it. Given our quantitative definition of an early-type galaxy based on bulge fraction and image smoothness, there are essentially two ways to transform late-type galaxies into early-type ones: 1) processes such as collisions and harassment that can fundamentally alter the structure of a galaxy by forming bulges and/or destroying disks and 2) quenching processes that can extinguish star forming regions responsible for some of the galaxy image asymmetries and also cause a fading of the disks. Applying the \\citet{poggianti06} scenario to our results, the \"threshold\" in $f_{et}$ values in our high redshift clusters (Figures~\\ref{vlt+sdss_efrac_sigma} and~\\ref{vlt+sdss_efrac_age}) could be explained by a population of primordial passive galaxies that formed at even higher redshifts. Most of our high redshift clusters have early-type fractions in the range 0.3-0.6 with no correlation with cluster velocity dispersion. Are these early-type fractions indeed consistent with a populations of primordial passive galaxies? Calculations done in \\citet{poggianti06} show that the fraction of galaxies at $z = 0.6$ that were present in haloes with masses greater than 3$\\times$10$^{12}$ M$_\\odot$ at $z$ = 2.5 is 0.4$\\pm$0.2. These primordial passive galaxies can therefore account for at least 2/3 (if not all) of the early-type populations in high redshift clusters, and their high formation redshift would explain the lack of dependence of $f_{et}$ on cluster velocity dispersion. One of our main results is that the early-type fractions of galaxy clusters increase from $z = 0.6 - 0.8$ to $z\\sim 0.08$ in clusters of all velocity dispersions. What kind of morphological transformation process(es) can lead to such an evolution? Collisions and harassment both depend on galaxy-galaxy interactions and the time a galaxy has spent within the cluster environment. Cluster velocity dispersion influences the number of interactions and their duration. Higher velocity dispersions in more massive clusters yield more interactions per unit time $N$ but with shorter durations $\\Delta t$ in a given time interval. One might therefore expect to see a peak in early-type type fraction at the cluster velocity dispersion where the product N$\\Delta t$ is maximized. No such peak is seen in our clusters. Ram-pressure stripping is expected to go as ($n_{ICM} v_{gal}^{2.4})/\\dot{M}_{rep}$ \\citep{gaetz87} with $n_{ICM}$, $v_{gal}$ and $\\dot{M}_{rep}$ being the density of the ICM, the velocity of the galaxies within the ICM and the rate at which galaxies can replenish their gas respectively. The fraction of passive galaxies should therefore be a relatively strong function of cluster velocity dispersion if quenching by ram pressure stripping is the dominant process. The number of post-starburst galaxies in EDisCS clusters does correlate with cluster velocity dispersion \\citep{poggianti09a}, but the uniform increase in early-type fractions at all cluster velocity dispersions observed going from EDisCS to SDSS clusters is not consistent with the intracluster medium being the main cause of the changes in cluster morphological content. Even though the early-type and [OII] emitter fractions in EDisCS and SDSS clusters show no correlation with cluster velocity dispersion \\citep[][and this work]{poggianti06}, there is a very strong correlation between $f_{et}$ and $f_{OII}$. This correlation is seen at both low and high cluster masses as well as at both low and high redshifts. Morphology and star formation therefore appear to be closely linked with one another over a wide range of environments and times. However, different structural transformation and quenching processes are thought to operate over different timescales \\citep[e.g.,][]{sanchez09}. Timescales range from 1-2 Gyr (based on typical cluster crossing times) for truncating star formation to 3-5 Gyr for totally extinguishing star formation in newly accreted galaxies \\citep{poggianti06,tonnesen09}. Looking at the evolution of EDisCS cluster red-sequence galaxies over 2 Gyr (from $z = 0.75$ to $z$ = 0.45), \\citet{sanchez09} found that morphological transformation and quenching of star formation indeed appeared to not be simultaneous. As noted in Section~\\ref{efrac-sigma}, the early-type fractions of mid-$z$ EDisCS clusters may be $\\sim$25$\\%$ higher than the ones of high-$z$ clusters. This change would therefore have taken place over a 2 Gyr interval in our adopted cosmology. However, the time baseline here between SDSS and EDisCS clusters is almost 6 Gyr, and, unfortunately, this is ample time to erase any difference arising from different timescales in the link between morphology and star formation. The lack of dependence of morphology and star formation on global cluster properties such as velocity dispersion raises the question of whether changes in galaxy properties are driven by more local effects or whether they occur outside of the cluster environment. Recent work \\citep{poggianti08,park09,bamford09,ellison09} have re-emphasized the strong link between galaxy properties and local galaxy density rather than cluster membership. Galaxy properties are seen to change at densities around 15-40 galaxies Mpc$^{-2}$ or projected separations of 20-30$h^{-1}$ kpc. Others \\citep[e.g.,][]{kautsch08,wilman09} have suggested that the galaxy group environment might be more conducive to galaxy transformation. Our observed evolution in early-type fraction as a function of redshift and the strong correlation between morphology and star formation at all cluster masses would support the idea that cluster membership is of lesser importance than other variables such as local density in determining galaxy properties. The properties of simulated clusters from the Millenium Simulation compare well with those of EDisCS and SDSS clusters. Their early-type fractions also show no dependence with cluster velocity dispersion in contrast to previous theoretical work \\citep[e.g.][]{diaferio01} but in agreement with observations. However, there is a definite lack of MS clusters with low early-type fractions at $z$ = 0 compared to the SDSS sample. It is important here to note that an early-type galaxy in the simulations was defined solely based on its bulge fraction because the simulations do not have the resolution required to model internal fine structures such as asymmetries. Given that real, early-type galaxies were also selected according to image smoothness, one would expect the early-type fractions of real clusters to be systematically lower. However, half of the SDSS clusters have low early-type fractions not seen in the simulations at $z$ = 0, and such a large discrepancy could only be explained by a significant population of real bulge-dominated galaxies with relatively large asymmetries. It is more likely that bulge formation in the simulations may be too efficient. The scatter in $f_{et}$ values for the simulated clusters with $\\sigma \\geq $ 600 km/s is also nearly three times smaller than observed in the real clusters (Section~\\ref{efrac-sigma}) which may indicate that the models may not include the right mixture of evolutionary processes at work on real galaxies. High-mass simulated clusters show a correlation between early-type fraction and star-forming fraction (albeit over narrower ranges than observed), but the correlation is not seen in the low-mass simulated clusters. This may be understood by high mass clusters having been formed long enough for evolutionary processes to have had enough time to act on galaxies to modify their properties whereas this is not necessarily the case for low-mass clusters. The fact that the correlation is observed in both low- and high-mass real clusters may be an indication that processes giving rise to the correlation may be more efficient (or altogether different) than modelled. It is also important to keep in mind here that the properties of a galaxy in these models are essentially driven by the mass of its parent halo." }, "0910/0910.4037_arXiv.txt": { "abstract": "{The seismic data obtained by CoRoT for the star HD~49933 enable us for the first time to measure \\emph{directly} the amplitudes and linewidths of solar-like oscillations for a star other than the Sun. From those measurements it is possible, as was done for the Sun, to constrain models of the excitation of acoustic modes by turbulent convection. } {We compare a stochastic excitation model described in Paper~I with the asteroseismology data for HD~49933, a star that is rather metal poor and significantly hotter than the Sun.} {Using the seismic determinations of the mode linewidths detected by CoRoT for HD~49933 and the theoretical mode excitation rates computed in Paper~I for the specific case of HD~49933, we derive the expected surface velocity amplitudes of the acoustic modes detected in HD~49933. Using a calibrated quasi-adiabatic approximation relating the mode amplitudes in intensity to those in velocity, we derive the expected values of the mode amplitude in intensity. } { Except at rather high frequency, our amplitude calculations are within 1-$\\sigma$ error bars of the mode surface velocity spectrum derived with the HARPS spectrograph. The same is found with respect to the mode amplitudes in intensity derived for HD~49933 from the CoRoT data. On the other hand, at high frequency ($\\nu \\gtrsim~1.9 $~mHz), our calculations depart significantly from the CoRoT and HARPS measurements. We show that assuming a solar metal abundance rather than the actual metal abundance of the star would result in a larger discrepancy with the seismic data. Furthermore, we present calculations which assume the ``new'' solar chemical mixture to be in better agreement with the seismic data than those that assumed the ``old'' solar chemical mixture.} { These results validate in the case of a star significantly hotter than the Sun and $\\alpha$~ Cen~A the main assumptions in the model of stochastic excitation. However, the discrepancies seen at high frequency highlight some deficiencies of the modelling, whose origin remains to be understood. We also show that it is important to take the surface metal abundance of the solar-like pulsators into account. } ", "introduction": "The amplitudes of solar-like oscillations result from a balance between excitation and damping. The mode linewidths are directly related to the mode damping rates. Once we can measure the mode linewidths, we can derive the theoretical value of the mode amplitudes from theoretical calculations of the mode excitation rates, which in turn can be compared to the available seismic constraints. This comparison allows us to test the model of stochastic mode excitation investigated in a companion paper \\citep[][hereafter Paper~I]{Samadi09a}. As shown in Paper~I, a moderate deficit of the surface metal abundance results in a significant decrease of the mode driving by turbulent convection. Indeed, by taking into account the measured iron-to-hydrogen abundance ([Fe/H]) of HD~49993 ([Fe/H]$=-0.37$), we have derived the theoretical values of the mode excitation rates ${\\cal P}$ expected for this star. The resulting value of ${\\cal P}$ is found to be about two times smaller than for a model with the same gravity and effective temperature, but with a solar metal abundance (i.e. [Fe/H]$=0$). The star HD~49933 was first observed in Doppler velocity by \\citet{Mosser05} with the HARPS spectrograph. More recently, this star has been observed twice by CoRoT. A first time this was done continuously during about 61 days (initial run, IR) and a second time continuously during about 137 days (first long run in the center direction, LRc01). The combined seismic analysis of these data \\citep{Benomar09b} has provided the mode linewidths as well as the amplitudes of the modes in intensity. Then, using mode linewidths obtained for HD 49933 with the CoRoT data and the theoretical mode excitation rates (obtained in Paper~I), we derive the expected values of the mode surface \\emph{velocity} amplitudes. We next compare these values with the mode velocity spectrum derived following \\citet{Kjeldsen05} with seismic data from the HARPS spectrograph \\citep{Mosser05}. Mode amplitudes in terms of \\emph{luminosity} fluctuations have also been derived from the CoRoT data for 17 radial orders. These data provide us with not only a constraint on the maximum of the mode amplitude but also with the frequency dependence. % The relative luminosity amplitudes $\\delta L/L$ are linearly related to the velocity amplitudes. This ratio is determined by the solution of the \\emph{non-adiabatic} pulsation equations and is independent of the stochastic excitation model \\citep[see][]{Houdek99}. Such a non-adiabatic calculation requires us to take into account, not only the radiative damping, but also the coupling between the pulsation and the turbulent convection. However, there are currently very significant uncertainties concerning the modeling of this coupling \\citep[for a recent review see ][]{Houdek08}. We relate further for the sake of simplicity the mode luminosity amplitudes to computed mode velocity amplitudes by assuming adiabatic oscillations as \\citet{Kjeldsen95}. Such a relation is calibrated in order to reproduce the helioseismic data. The comparison between theoretical values of the mode amplitudes (both in terms of surface velocity and intensity) constitutes a test of the stochastic excitation model with a star significantly different from the Sun and {\\acenA}. In addition it is also possible to test the validity of the calibrated quasi-adiabatic relation, since both mode amplitudes, in terms of surface velocity and intensity, are available for this star, This paper is organized as follows: We describe in Sect.~\\ref{velocity} the way mode amplitudes in terms of surface velocity $v_s$ are derived from the theoretical values of ${\\cal P}$ and from the measured mode linewidths ($\\Gamma$). Then, we compare the theoretical values of the mode surface velocity with the seismic constraint obtained from HARPS observations. We describe in Sect.~\\ref{intensity} the way mode amplitudes in terms of intensity fluctuations $\\delta L/L$ are derived from theoretical values of $v_s$ and compare $\\delta L/L$ with the seismic constraints obtained from the CoRoT observations. Finally, Sects. \\ref{Discussion} and \\ref{Conclusion} are dedicated to a discussion and conclusion respectively. ", "conclusions": "\\label{Conclusion} From the mode linewidths measured by CoRoT and theoretical mode excitation rates derived for HD~49933, we have derived the expected mode surface velocities $v_s$ which we have compared with $v_{\\rm HARPS}$, the mode velocity spectrum derived from the seismic observations obtained with the HARPS spectrograph \\citep{Mosser05}. Except at high frequency ($\\nu \\gtrsim$~1.9~mHz), the agreement between computed $v_s$ and $v_{\\rm HARPS}$ is within the 1-$\\sigma$ domain associated with the seismic data from the HARPS spectrograph. However, there is a clear tendency to overestimate $v_{\\rm HARPS}$ above $\\nu~\\sim$~1.9~mHz,. Using a \\emph{calibrated} quasi-adiabatic approximation to relate the mode velocity to the mode amplitude in intensity (Eq.~\\ref{dL_Vad}), we have derived for the case of HD~49933 the expected mode amplitudes in intensity. Computed mode intensity fluctuations, $\\delta L/L$, are within 1-$\\sigma$ in agreement with the seismic constraints derived from the CoRoT data \\citep{Benomar09b}. However, as for the velocity, there is a clear tendency at high frequency ($\\nu \\gtrsim$~1.9~mHz) towards over-estimated $\\delta L/L$ compared to the CoRoT observations. Calculations that assume a solar surface metal abundance result, both in velocity and in intensity, in amplitudes larger by $\\sim$~35\\,\\% around the peak frequency ($\\nu_{\\rm max} \\simeq$ 1.8~mHz) and by up to a factor of two at lower frequency. It follows that, ignoring the current surface metal abundance of the star results in a more severe over-estimation of the computed amplitudes compared with observations. This illustrates the importance of taking the surface metal abundance of the solar-like pulsators into account when modeling the mode driving. In addition, we point out that the \\citet{GN93} solar chemical mixture results in mode amplitudes larger by about 15\\,\\% with respect to calculations that assume the ''new'' solar abundance by \\citet{Asplund05}. However, this increase remains significantly smaller than the uncertainties associated with current seismic measurements. Since both mode amplitudes in terms of surface velocity and intensity are available for this star, it was possible to test the validity of the calibrated quasi-adiabatic relation (\\eq{dL_Vad}). Our comparison shows that this relation provides the correct scaling, at least at the level of the present seismic precisions, . Both in terms of surface velocity and of intensity, the differences between predicted and observed mode amplitudes are within the 1-$\\sigma$ uncertainty domain, except at high frequency. This result then validates for low frequency modes the basic underlying physical assumptions included in the theoretical model of stochastic excitation for a star significantly different in effective temperature, surface gravity, turbulent Mach number ($M_t$) and metallicity compared to the Sun or $\\alpha$~Cen~A. As discussed in Sect.~\\ref{Discussion}, the clear discrepancy between predicted and observed mode amplitudes seen at high frequency may have two possible origins: First, a canceling between the entropy contribution and the Reynolds stress is expected to occur and to be important around and above the frequency of the maximum of the mode excitation rates (see Sect.~\\ref{canceling_entropy_reynolds}). Second, the assumption called the ``scale length separation'' \\citep{Samadi08b} may also result in an over-estimation of the mode amplitudes at high frequency (see Sect.~\\ref{scale length separation}). These issues will be investigated in a forthcoming paper." }, "0910/0910.3421_arXiv.txt": { "abstract": "\\vsco\\ is an eclipsing dwarf nova that had attracted little attention from X-ray astronomers until it was proposed as the identification of an \\xte\\ all-sky slew survey (XSS) source. Here we report on the pointed X-ray observations of this object using \\suzaku. \\vsco\\ was in quiescence at the time, as indicated by the coordinated optical photometry we obtained at the South African Astronomical Observatory. Our \\suzaku\\ data show \\vsco\\ to be X-ray bright, with a highly absorbed spectrum. Most importantly, we have discovered a partial X-ray eclipse in \\vsco. This is the first time that a partial eclipse is seen in X-ray light curves of a dwarf nova. Our preliminary simulations demonstrate that the partial X-ray eclipse can be in principle reproduced if the white dwarf in \\vsco\\ is partially eclipsed. Higher quality observations of this object have the potential to place significant constraints on the latitudinal extent of the X-ray emission region and thereby discriminating between an equatorial boundary layer and a spherical corona. The partial X-ray eclipse therefore makes \\vsco\\ a key object in understanding the physics of accretion in quiescent dwarf nova. ", "introduction": "Cataclysmic variables (CVs), in which a white dwarf primary accretes from a Roche-lobe filling, late-type secondary (see \\citealt{tome} for a comprehensive review), are an excellent laboratory for the physics of accretion. In the subclass of dwarf novae, accretion proceeds via a disk that switches between the low (quiescence) and high (outburst) luminosity states. The visible light of a quiescent dwarf nova is usually dominated by the bright spot, where the accretion flow from the secondary hits the outer edge of the disk, as well as the photosphere of the white dwarf. In outburst, the disk becomes the dominant source of visible light. In contrast, X-ray observations of dwarf novae probe the accretion flow in the immediate vicinity of the white dwarf, such as an optically thin boundary layer \\citep{PR1985a}, since the accretion disk in a dwarf nova is too cool to emit X-rays. Although the disk instability model is a highly successful framework for explaining the dwarf nova outbursts, there are details that defy the prediction of the basic version of the model \\citep{L2001}. In particular, the observed X-ray luminosities of quiescent dwarf nova (often of order 10$^{31}$ \\eps; \\citealt{Bea2005}) imply an accretion rate onto the white dwarf that is much higher than predicted. Proposed modifications of the disk instability model include the coronal siphon flow \\citep{MM1994}, which might lead to accretion over a much wider area of the white dwarf surface than through a boundary layer, and a weakly magnetic white dwarf \\citep{LP1992}. In the latter model, the magnetic field is too weak to control the accretion flow during outburst, but strong enough to do so during quiescence. \\vsco\\ is an eclipsing \\citep{BSG2000} dwarf nova with an orbital period of 1.8 hr \\citep{T1999}. According to \\cite{KMU2002}, \\vsco\\ has a quiescent magnitude of $\\sim$14.5, and has an outburst every $\\sim$30 days during which it reaches magnitude $\\sim$12.5. Extensive photometry by Warner and collaborators \\citep{WWP2003,PWW2006} have revealed quasi-periodic oscillations and dwarf nova oscillations in \\vsco, but a strictly periodic signal was never found. Therefore, existing optical data point strongly towards a dwarf nova classification and argue against an intermediate polar (IP, or DQ Her type systems; \\citealt{P1994}) classification. IPs are a subset of magnetic CVs in which the primary's magnetic field disrupts the inner accretion disk, channeling the flow to the magnetic polar region(s); the spin period of the magnetic white dwarf is a strict clock that characterize IPs. Thus, \\vsco\\ joins a growing number of eclipsing dwarf novae below the period gap. Citing the variable shape of the eclipses and the relatively small eclipse amplitude (often less than 0.75 mag), \\cite{BSG2000} argue for a grazing eclipse of the bright spot and the disk, but not of the white dwarf. On the other hand, \\cite{Mea2000} argue that the white dwarf is eclipsed, based on the fact that the spectroscopic conjunction of the disk (the red-to-blue crossing of the emission line radial velocities) occurs at mid-eclipse. From the radial velocity curves, they also infer a white dwarf mass of $\\sim$0.5--0.6 M$_\\odot$ and a mass ratio of $\\sim$0.2--0.3. However, \\cite{Mea2001} prefer a higher mass (0.89 M$_\\odot$) white dwarf and an inclination angle of 72.5 degrees based on their analysis of the emission line radial velocities, although this value depends in part on the assumed mass-radius relationship for the secondary. Despite the current lack of a consensus regarding the system parameters, the eclipsing nature of \\vsco\\ makes it an important target for detailed studies. Moreover, it is a nearby system with parallax-estimated distance of 155$^{+58}_{-34}$ pc \\citep{T2003}. Although the \\rosat\\ detection was already noted by \\cite{Kea1998}, \\vsco\\ did not draw the attention of X-ray astronomers until it was listed among the \\xte\\ all-sky slew survey (XSS) sources \\citep{Rea2004}, along with three other dwarf novae (SS~Aur, V426~Oph, and SU~UMa). The estimated luminosities of these 4 systems are all just under 10$^{32}$ \\eps\\ in the 2--10 keV band, placing them near the upper end for non-magnetic CVs ($10^{30}$--$3 \\times 10^{32}$ in 0.1--100 keV for the \\asca\\ sample; \\citealt{Bea2005}). However, the XSS is based on the data taken during slews of the non-imaging \\xte\\ PCA detector. For relatively faint sources such as these dwarf novae, the positional errors are considerable, and misidentification is a possibility. Hence it is important to confirm the proposed identification of XSS dwarf novae. This is particularly true for \\vsco, which had never been the subject of a pointed X-ray observation above 2 keV. Although \\vsco\\ is securely detected as a \\rosat\\ all-sky survey (RASS) source, the ratio of XSS (2.38 c/s in the 3--8 keV band) to RASS (0.35) count rates is high, compared to, e.g., SS~Aur which has the same RASS count rate but was detected at 0.75 c/s in the XSS 3--8 keV band. This leaves open the possibility that \\vsco\\ is only a partial contributor to the XSS flux. For this reason and also because of the potential return in studying the X-ray properties of this eclipsing dwarf nova, we performed a pointed observation using \\suzaku, along with contemporaneous optical photometry. ", "conclusions": "We performed \\suzaku\\ observations of the eclipsing dwarf nova, \\vsco. We confirm that it is an X-ray luminous dwarf nova, and moreover, report the discovery of a partial X-ray eclipse. In the past, both optical and X-ray observers have concentrated on systems that exhibit a total eclipse of the white dwarf. They provide a strong constraint on the height of the X-ray emission region above the white dwarf. However, eclipse light curves are one-dimensional, and one can only derive one-dimensional constraints from the eclipse analysis. The X-ray emission region, on the other hand, has two dimensions (height and latitudinal extent) assuming azimuthal symmetry. It is therefore necessary to observe an ensemble of eclipsing dwarf novae, in which the limb of the secondary cuts across the emission region from different angles, to be able to constrain the height and the latitudinal extent simultaneously. Inclusion of a partially eclipsing system in such a study will be a huge step forward in our quest to constrain the X-ray emission region size in two dimensions. Future X-ray observations of \\vsco, together with improved estimates of system parameters from other wavelengths, have the potential to test the boundary layer picture of X-ray emission in quiescent dwarf novae, and furthermore to constrain the latitudinal extent of such a boundary layer." }, "0910/0910.4806_arXiv.txt": { "abstract": "We present a careful and detailed light curve analysis of RR Lyrae stars in the Small Magellanic Cloud (SMC) discovered by the Optical Gravitational Lensing Experiment (OGLE) project. Out of 536 single mode RR Lyrae stars selected from the database, we have investigated the physical properties of 335 `normal looking' RRab stars and 17 RRc stars that have good quality photometric light curves. We have also been able to estimate the distance modulus of the cloud which is in good agreement with those determined from other independent methods. The Fourier decomposition method has been used to study the basic properties of these variables. Accurate Fourier decomposition parameters of 536 RR Lyrae stars in the OGLE-II database are computed. Empirical relations between the Fourier parameters and some physical parameters of these variables have been used to estimate the physical parameters for the stars from the Fourier analysis. Further, the Fourier decomposition of the light curves of the SMC RR Lyrae stars yields their mean physical parameters as: [Fe/H] = -1.56\\,$\\pm\\,0.25$, M = 0.55 $\\pm $\\,0.01\\,M$_{\\odot}$, T$_{\\rm eff} = 6404\\, \\pm 12$ K, $\\log \\rm L = 1.60\\,\\pm0.01\\,\\rm L_\\odot$ and M$\\rm _V = 0.78\\,\\pm0.02 $ for 335 RRab variables and [Fe/H] = -1.90\\,$\\pm$\\, 0.13, M = 0.82 $\\pm $\\,0.18\\,M$_{\\odot}$, T$_{\\rm eff} = 7177\\,\\pm 16$ K, $\\log\\,\\rm L = 1.62\\,\\pm \\,0.02\\,\\rm L_{\\odot}$ and M$\\rm _V = 0.76\\,\\pm 0.05$ for 17 RRc stars. Using the absolute magnitude together with the mean magnitude, intensity-weighted mean magnitude and the phase-weighted mean magnitude of the RR Lyrae stars, the mean distance modulus to the SMC is estimated to be 18.86\\,$\\pm$0.01 mag, 18.83\\,$\\pm$0.01 mag and 18.84\\,$\\pm$0.01 mag respectively from the RRab stars. From the RRc stars, the corresponding distance modulus values are found to be 18.92\\,$\\pm$0.04 mag, 18.89\\,$\\pm$0.04 mag and 18.89\\,$\\pm$0.04 mag respectively. Since Fourier analysis is a very powerful tool for the study of the physical properties of the RR Lyrae stars, we emphasize the importance of exploring the reliability of the calculation of Fourier parameters together with the uncertainty estimates keeping in view the large collections of photometric light curves that will become available from variable star projects of the future. ", "introduction": "RR Lyrae (RRL) stars have played an important role in determining the cosmic distance scale in modern astronomy. While several studies have used classical Cepheids as primary distance indicators, RRL stars have received substantial attention in solving the distance scale problem. Apart from distance determinations, RRL stars are of particular importance as a test bed for the theories of stellar and galactic structure and evolution. They are the tracers of old stellar populations of the bulge, disk and halo components that are present everywhere. RRL stars are radially pulsating A-F variable stars with periods in the 0.2 - 1.2 day range, and amplitude of variation $ \\leq 2$ mag. They can be easily identified, and play a key role as the cornerstone of the Population II distance scale. They are extensively used to determine distances to old and sufficiently metal-poor systems, where they are commonly found in large numbers. In particular, RRL stars are present in globular clusters (GCs) and the dwarf galaxies in the neighborhood of the Milky Way (Greco et al. 2007), and have also been identified in the M31 field (Brown et al. 2004, Dolphin et al. 2004), in some M31 companions (Pritzl et al. 2005), and in at least four M31 GCs (Clementini et al. 2001). Distances to the Large Magellanic Cloud (LMC) for the population II objects are based on the luminosity of the RRL stars (Clementini et al. 2002). There have been a very few studies to estimate the distance scale of the SMC using the RRL stars. Using four RRL light curves of the SMC cluster NGC 121, Walker \\& Mack (1988) have estimated the distance modulus of the cluster to be 18.86$\\pm$0.07 mag. On the other hand, using 22 RR Lyrae stars surveyed around 1.3 square degrees near the northeast arm of the SMC field NGC 361, Smith et al. (1992) obtained the distance modulus of the SMC as 18.90$\\pm$0.16 mag. Also, there are other studies of estimating the distance scale of the SMC based on the binary star light curves and double mode Cepheids. Harries et al. (2003) reported that the distance modulus to the SMC is of the order of 18.89$\\pm$0.14 mag taking 10 eclipsing binaries in the SMC. By selecting 40 eclipsing binaries of spectral type O and B in the SMC, Hilditch et al. (2005) have derived the fundamental parameters of the binaries and refined the distance modulus to the SMC to 18.91$\\pm$0.1 mag. Also Kov\\'{a}cs et al. (2000) found the distance modulus of the SMC to be 19.05$\\pm$0.017 mag based on the photometric data of double-mode Cepheids from the OGLE project. The stellar atmospheric parameters of effective temperature (T$_{\\rm eff}$) and surface gravity (log g) are of fundamental astrophysical importance. They are the prerequisites to any detailed abundance analysis and define the physical conditions in the stellar atmosphere and hence are directly related to the physical properties mass (M), radius (R) and luminosity (L) of the star. In this paper, we present an independent analysis of 536 SMC RRL stars discovered by the OGLE project (Soszy\\'{n}sky et al. 2002). The OGLE database is a very wealthy resource for studying the characteristics of variable stars in the Galaxy, LMC and SMC. For the first time, we make use of the OGLE SMC data to estimate the distance scale of the SMC using a large number of well-sampled RRL light curves. The road map of the present investigation is to perform a Fourier analysis of the RRL stars in order to estimate their physical parameters and hence the SMC distance scale. We employ the Fourier decomposition technique which is used extensively to characterize the observed photometric light curves of RRL and other types of variables. The accurate determination of the Fourier coefficients is, therefore, an important task. We have performed an independent automated Fourier analysis of all the RRL light curves selected in this paper by a computer code developed by us. In section 2 we give a brief description of the database that we use and the procedure of removing the outliers from the light curve data. We present Fourier decomposition of the light curves in section 3. We also describe the use of the unit-lag auto-correlation function for finding out the optimal order of the fit to the RRL light curves. Section 4 describes the error analysis of the Fourier decomposition parameters $\\rm \\phi_{i1}$ and R$_{i1}$. Section 5 describes the calibration of the I band data to the V band. In section 6, we describe the various physical parameters of the RRLs obtained by using empirical relations from the literature. Section 7 describes the distance determination of the SMC. Lastly, in section 8, we present the conclusions of our study. ", "conclusions": "In this paper we have derived the physical parameters of 352 RR Lyrae stars (335 RRab and 17 RRc) of the SMC from OGLE-II I band database using the Fourier decomposition of their light curves. The stars were selected based on the quality of their light curves. I band Fourier coefficients have been converted to V band using the inter-relations obtained from the observational data of B06 and W94. Using the Fourier decomposition method, we find the mean physical parameters: [Fe/H] = -1.56\\,$\\pm\\,0.25$, M = 0.55 $\\pm $\\,0.01\\,M$_{\\odot}$, T$_{\\rm eff} = 6404\\, \\pm 12$ K, $\\log \\rm L = 1.60\\,\\pm0.01\\,\\rm L_\\odot$ and M$\\rm _V = 0.78\\,\\pm0.02 $ for 335 RRab variables and [Fe/H] = -1.90\\,$\\pm$\\, 0.13, M = 0.82 $\\pm $\\,0.18\\,M$_{\\odot}$, T$_{\\rm eff} = 7177\\,\\pm 16$ K, $\\log\\,\\rm L = 1.62\\,\\pm \\,0.02\\,\\rm L_{\\odot}$ and M$\\rm _V = 0.76\\,\\pm 0.05$ for 17 RRc stars. Mean distance modulus to the SMC was calculated from the 352 light curves by using mean magnitude, intensity weighted mean magnitude and phase weighted mean magnitude. The values of the distance modulus are found to be in good agreement with independent studies. Locations of the RRL stars in the H-R diagram clearly show that the estimates of the parameters determined from the Fourier decomposition method are consistent with the theoretical blue and red edges of the instability strip calculated by Bono et al. (1995). The calculations of the radii of the RRLs by using the Fourier decomposition technique are in agreement with the theoretical period-radius-metallicity relations of Marconi et al. (2005) obtained from a completely different approach of non-linear convective modelling. \\begin{figure} \\centering \\includegraphics[height=8cm,width=9cm]{fig17.eps} \\vspace{10pt} \\caption{Histogram of the metallicity of the RRLs used for the analysis. The partially filled histogram shows the metallicity distribution of the 335 RRab stars, whereas the fully filled histogram shows the metallicity distribution of the 17 RRc stars.} \\label{Fig 14}% \\end{figure} \\begin{table*} \\caption{Physical parameters extracted from Fourier coefficients for 335 RRab variables. Errors represent the uncertainties in the Fourier parameters.} \\begin{center} \\scalebox{0.9}{ \\begin{tabular}{ccccccccccccccccc} \\\\ \\hline \\hline OGLE ID &M$_{V}$&\\rm $\\log$\\,(L/L$_{\\odot})$&[\\rm Fe/H]&T$_{\\rm eff}$&\\rm M/M$_{\\odot}$&\\rm log{\\rm g}&$\\rm R/R_{\\odot} (FD)$& $\\rm R/R_{\\odot} $ (Mar) \\\\ ~~~~~~~(1)&(2)&(3)&(4)&(5)&(6)&(7)&(8)&(9)\\\\ \\hline \\hline 004801.59-733021.5& 0.93$\\pm$ 0.02& 1.53$\\pm$ 0.01& -0.91$\\pm$ 0.19& 6751$\\pm$ 9& 0.61$\\pm$ 0.01& 2.97$\\pm$ 0.02& 4.27$\\pm$ 0.02& 4.31$\\pm$ 0.03\\\\ 005300.26-725136.6& 0.92$\\pm$ 0.02& 1.54$\\pm$ 0.01& -1.18$\\pm$ 0.19& 6702$\\pm$ 9& 0.64$\\pm$ 0.01& 2.97$\\pm$ 0.02& 4.40$\\pm$ 0.02& 4.43$\\pm$ 0.03\\\\ 003841.21-734422.9& 0.90$\\pm$ 0.01& 1.54$\\pm$ 0.01& -1.00$\\pm$ 0.18& 6758$\\pm$ 9& 0.60$\\pm$ 0.01& 2.95$\\pm$ 0.02& 4.33$\\pm$ 0.02& 4.41$\\pm$ 0.03\\\\ 003727.78-731454.9& 0.92$\\pm$ 0.01& 1.53$\\pm$ 0.01& -0.88$\\pm$ 0.16& 6755$\\pm$ 8& 0.59$\\pm$ 0.01& 2.95$\\pm$ 0.01& 4.29$\\pm$ 0.02& 4.38$\\pm$ 0.02\\\\ 005728.85-723454.6& 0.93$\\pm$ 0.02& 1.52$\\pm$ 0.01& -0.76$\\pm$ 0.21& 6739$\\pm$10& 0.57$\\pm$ 0.01& 2.94$\\pm$ 0.02& 4.26$\\pm$ 0.02& 4.36$\\pm$ 0.03\\\\ 005728.85-723454.6& 0.93$\\pm$ 0.02& 1.52$\\pm$ 0.01& -0.76$\\pm$ 0.21& 6739$\\pm$10& 0.57$\\pm$ 0.01& 2.94$\\pm$ 0.02& 4.26$\\pm$ 0.02& 4.36$\\pm$ 0.03\\\\ 005026.32-732418.2& 0.94$\\pm$ 0.02& 1.52$\\pm$ 0.01& -1.00$\\pm$ 0.20& 6653$\\pm$ 9& 0.60$\\pm$ 0.01& 2.94$\\pm$ 0.02& 4.39$\\pm$ 0.02& 4.48$\\pm$ 0.03\\\\ 004639.18-731324.7& 0.89$\\pm$ 0.02& 1.54$\\pm$ 0.01& -0.81$\\pm$ 0.20& 6800$\\pm$10& 0.56$\\pm$ 0.01& 2.93$\\pm$ 0.02& 4.29$\\pm$ 0.02& 4.42$\\pm$ 0.03\\\\ 003816.46-732449.2& 0.92$\\pm$ 0.02& 1.52$\\pm$ 0.01& -0.74$\\pm$ 0.19& 6738$\\pm$ 9& 0.55$\\pm$ 0.01& 2.93$\\pm$ 0.02& 4.29$\\pm$ 0.02& 4.42$\\pm$ 0.03\\\\ 005110.48-730750.0& 0.87$\\pm$ 0.01& 1.55$\\pm$ 0.01& -0.85$\\pm$ 0.18& 6805$\\pm$ 9& 0.55$\\pm$ 0.01& 2.92$\\pm$ 0.02& 4.33$\\pm$ 0.02& 4.48$\\pm$ 0.03\\\\ 010452.90-724025.9& 0.95$\\pm$ 0.01& 1.52$\\pm$ 0.01& -1.00$\\pm$ 0.19& 6599$\\pm$ 9& 0.59$\\pm$ 0.01& 2.92$\\pm$ 0.02& 4.44$\\pm$ 0.02& 4.55$\\pm$ 0.03\\\\ 005646.16-723452.2& 0.88$\\pm$ 0.01& 1.55$\\pm$ 0.01& -1.06$\\pm$ 0.18& 6691$\\pm$ 9& 0.57$\\pm$ 0.01& 2.90$\\pm$ 0.02& 4.48$\\pm$ 0.02& 4.65$\\pm$ 0.03\\\\ 005957.83-730647.6& 0.87$\\pm$ 0.01& 1.55$\\pm$ 0.01& -0.99$\\pm$ 0.17& 6717$\\pm$ 8& 0.56$\\pm$ 0.01& 2.90$\\pm$ 0.01& 4.44$\\pm$ 0.02& 4.63$\\pm$ 0.03\\\\ 005458.09-724948.9& 0.89$\\pm$ 0.01& 1.55$\\pm$ 0.00& -1.11$\\pm$ 0.15& 6646$\\pm$ 7& 0.58$\\pm$ 0.01& 2.90$\\pm$ 0.01& 4.51$\\pm$ 0.01& 4.68$\\pm$ 0.02\\\\ 004758.98-732241.3& 0.89$\\pm$ 0.02& 1.54$\\pm$ 0.01& -0.95$\\pm$ 0.26& 6683$\\pm$12& 0.56$\\pm$ 0.01& 2.90$\\pm$ 0.02& 4.45$\\pm$ 0.03& 4.63$\\pm$ 0.04\\\\ 010535.93-720621.6& 0.94$\\pm$ 0.02& 1.53$\\pm$ 0.01& -0.98$\\pm$ 0.30& 6583$\\pm$13& 0.57$\\pm$ 0.01& 2.89$\\pm$ 0.03& 4.49$\\pm$ 0.03& 4.67$\\pm$ 0.05\\\\ 010516.55-722526.5& 0.90$\\pm$ 0.02& 1.53$\\pm$ 0.01& -0.76$\\pm$ 0.20& 6702$\\pm$ 9& 0.53$\\pm$ 0.01& 2.89$\\pm$ 0.02& 4.38$\\pm$ 0.02& 4.59$\\pm$ 0.03\\\\ 004721.26-731135.5& 0.89$\\pm$ 0.01& 1.55$\\pm$ 0.01& -1.02$\\pm$ 0.17& 6640$\\pm$ 8& 0.56$\\pm$ 0.01& 2.88$\\pm$ 0.02& 4.52$\\pm$ 0.02& 4.73$\\pm$ 0.03\\\\ 005504.67-731106.4& 0.91$\\pm$ 0.01& 1.54$\\pm$ 0.01& -1.08$\\pm$ 0.19& 6583$\\pm$ 9& 0.57$\\pm$ 0.01& 2.88$\\pm$ 0.02& 4.56$\\pm$ 0.02& 4.76$\\pm$ 0.03\\\\ 004306.71-733527.9& 0.92$\\pm$ 0.02& 1.53$\\pm$ 0.01& -0.89$\\pm$ 0.20& 6605$\\pm$10& 0.54$\\pm$ 0.01& 2.88$\\pm$ 0.02& 4.49$\\pm$ 0.02& 4.70$\\pm$ 0.03\\\\ \\hline \\hline \\end{tabular} } \\end{center} Complete table is available in the electronic form. \\end{table*} \\begin{table*} \\caption{Physical parameters extracted from Fourier coefficients for 17 RRc variables. Errors represent the uncertainties in the Fourier parameters. } \\scalebox{0.9}{ \\begin{tabular}{cccccccccc} \\\\ \\hline \\hline OGLE ID &\\rm M$_{\\rm V}$(K98)&$\\rm \\log (\\rm L/L_{\\odot}) (SC93) $&[\\rm Fe/H]&T$_{\\rm eff}$&\\rm M/M$_{\\odot}$&\\rm log \\rm g&$\\rm R/R_{\\odot} (FD)$&$\\rm R/R_{\\odot}$ (Mar) \\\\ ~~~~~~~(1)&(2)&(3)&(4)&(5)&(6)&(7)&(8)&(9) \\\\ \\hline \\hline 005451.72-723850.4& 0.85$\\pm$ 0.03& 1.66$\\pm$ 0.02& -1.10$\\pm$ 0.17& 7457$\\pm$17& 0.59$\\pm$ 0.19& 3.00$\\pm$ 0.33& 4.07$\\pm$ 0.03& 4.19$\\pm$ 0.02\\\\ 005115.64-724739.2& 0.78$\\pm$ 0.04& 1.67$\\pm$ 0.02& -1.21$\\pm$ 0.16& 7425$\\pm$17& 0.62$\\pm$ 0.19& 3.00$\\pm$ 0.31& 4.19$\\pm$ 0.04& 4.25$\\pm$ 0.02\\\\ 010245.99-721132.7& 0.83$\\pm$ 0.04& 1.83$\\pm$ 0.03& -2.02$\\pm$ 0.11& 7165$\\pm$18& 1.13$\\pm$ 0.07& 3.04$\\pm$ 0.06& 5.35$\\pm$ 0.05& 4.63$\\pm$ 0.02\\\\ 004631.02-724658.8& 0.78$\\pm$ 0.04& 1.75$\\pm$ 0.03& -1.67$\\pm$ 0.17& 7266$\\pm$20& 0.77$\\pm$ 0.18& 2.97$\\pm$ 0.24& 4.79$\\pm$ 0.05& 4.65$\\pm$ 0.02\\\\ 004902.13-724513.6& 0.82$\\pm$ 0.03& 1.76$\\pm$ 0.02& -1.73$\\pm$ 0.13& 7247$\\pm$15& 0.80$\\pm$ 0.18& 2.97$\\pm$ 0.23& 4.87$\\pm$ 0.04& 4.69$\\pm$ 0.02\\\\ 010453.59-723756.0& 0.78$\\pm$ 0.03& 1.83$\\pm$ 0.01& -2.11$\\pm$ 0.08& 7115$\\pm$10& 0.93$\\pm$ 0.17& 2.94$\\pm$ 0.18& 5.44$\\pm$ 0.04& 5.09$\\pm$ 0.01\\\\ 010029.57-725454.2& 0.77$\\pm$ 0.04& 1.81$\\pm$ 0.02& -2.03$\\pm$ 0.09& 7136$\\pm$10& 0.85$\\pm$ 0.19& 2.93$\\pm$ 0.22& 5.32$\\pm$ 0.05& 5.10$\\pm$ 0.01\\\\ 010029.57-725454.2& 0.77$\\pm$ 0.04& 1.81$\\pm$ 0.02& -2.03$\\pm$ 0.09& 7136$\\pm$10& 0.85$\\pm$ 0.19& 2.93$\\pm$ 0.22& 5.32$\\pm$ 0.05& 5.10$\\pm$ 0.01\\\\ 004327.25-724542.4& 0.76$\\pm$ 0.03& 1.75$\\pm$ 0.02& -1.62$\\pm$ 0.13& 7238$\\pm$13& 0.64$\\pm$ 0.22& 2.89$\\pm$ 0.34& 4.80$\\pm$ 0.04& 4.96$\\pm$ 0.02\\\\ 010231.17-722134.0& 0.69$\\pm$ 0.04& 1.83$\\pm$ 0.03& -2.13$\\pm$ 0.15& 7107$\\pm$18& 0.91$\\pm$ 0.18& 2.93$\\pm$ 0.19& 5.46$\\pm$ 0.06& 5.16$\\pm$ 0.02\\\\ 005556.74-732133.6& 0.77$\\pm$ 0.05& 1.77$\\pm$ 0.03& -1.80$\\pm$ 0.17& 7193$\\pm$18& 0.70$\\pm$ 0.21& 2.89$\\pm$ 0.31& 5.00$\\pm$ 0.06& 5.07$\\pm$ 0.03\\\\ 003934.35-730433.9& 0.76$\\pm$ 0.03& 1.79$\\pm$ 0.02& -1.92$\\pm$ 0.13& 7161$\\pm$13& 0.76$\\pm$ 0.21& 2.90$\\pm$ 0.28& 5.16$\\pm$ 0.04& 5.13$\\pm$ 0.02\\\\ 010647.29-723053.7& 0.73$\\pm$ 0.04& 1.81$\\pm$ 0.02& -2.03$\\pm$ 0.13& 7124$\\pm$14& 0.77$\\pm$ 0.21& 2.88$\\pm$ 0.27& 5.31$\\pm$ 0.04& 5.27$\\pm$ 0.02\\\\ 005155.37-724909.5& 0.73$\\pm$ 0.04& 1.77$\\pm$ 0.02& -1.78$\\pm$ 0.17& 7183$\\pm$16& 0.65$\\pm$ 0.23& 2.86$\\pm$ 0.35& 5.00$\\pm$ 0.05& 5.19$\\pm$ 0.03\\\\ 010316.37-724816.2& 0.66$\\pm$ 0.12& 1.90$\\pm$ 0.03& -2.56$\\pm$ 0.13& 6968$\\pm$18& 1.14$\\pm$ 0.12& 2.92$\\pm$ 0.11& 6.20$\\pm$ 0.15& 5.54$\\pm$ 0.02\\\\ 010316.37-724816.2& 0.66$\\pm$ 0.12& 1.90$\\pm$ 0.03& -2.56$\\pm$ 0.13& 6968$\\pm$18& 1.14$\\pm$ 0.12& 2.92$\\pm$ 0.11& 6.20$\\pm$ 0.15& 5.54$\\pm$ 0.02\\\\ 005527.97-724136.8& 0.76$\\pm$ 0.04& 1.81$\\pm$ 0.03& -2.03$\\pm$ 0.16& 7118$\\pm$17& 0.75$\\pm$ 0.22& 2.87$\\pm$ 0.29& 5.32$\\pm$ 0.05& 5.35$\\pm$ 0.03\\\\ \\hline \\hline \\end{tabular} } \\end{table*}" }, "0910/0910.3751_arXiv.txt": { "abstract": "Various theoretical and observational results have been reported regarding the presence/absence of net electric currents in the sunspots. The limited spatial resolution of the earlier observations perhaps obscured the conclusions. We have analyzed 12 sunspots observed from Hinode (SOT/SP) to clarify the issue. The azimuthal and radial components of magnetic fields and currents have been derived. The azimuthal component of the magnetic field of sunspots is found to vary in sign with azimuth. The radial component of the field also varies in magnitude with azimuth. While the latter pattern is a confirmation of the interlocking combed structure of penumbral filaments, the former pattern shows that the penumbra is made up of a ``curly interlocking combed\" magnetic field. The azimuthally averaged azimuthal component is seen to decline much faster than 1/$\\varpi$ in the penumbra, after an initial increase in the umbra, for all the spots studied. This confirms the confinement of magnetic fields and absence of a net current for sunspots as postulated by \\cite{parker96}. The existence of a global twist for a sunspot even in the absence of a net current is consistent with a fibril-bundle structure of the sunspot magnetic fields. ", "introduction": "Sunspots have shown evidence for twist even from the time of \\cite{hale25,hale27} who postulated the hemispheric rule for the chirality of chromospheric whirls. This was later confirmed with a larger data set by \\cite{rich41}. Evidence for photospheric chirality could be seen in early continuum images of sunspots, obtained with exceptional image quality. Later, photospheric vector magnetograms showed global twist inferred from the non-vanishing averages of the force-free parameter (\\cite{pcm94,hagi04,nand06} and references therein). The non-force-free nature of photospheric magnetic field in the sunspots, prompted \\cite{tiw09a} to propose the signed shear angle (SSA) as a more robust measure of the global twist of the sunspot magnetic field. Although, the sign of SSA matches well with the sign of the global alpha parameter, the magnitudes are not so well correlated. The physical significance of a globally averaged $\\alpha$ parameter rests heavily on the existence of a net current in the photospheric sunspot magnetic field. One way of arriving at a global $\\alpha$ is by taking the ratio of total vertical current to the total flux (integral method). This value was found to agree with the values obtained by other methods \\citep{hagi04}. For a monolithic sunspot magnetic field, the global twist and net current is expected to be well correlated by Ampere's Law. However, the existence of a net current is ruled out theoretically for fibril bundles as well as for monolithic fields with azimuthal field decreasing faster than 1/$\\varpi$, where $\\varpi$ is the radial distance from the spot center \\citep{parker96}. Several attempts to resolve this problem using vector magnetograms have not been very conclusive so far \\citep{wilk92,leka96,wheat2000}. A resolution of this problem can be used to disentangle the relation between global twist and the global $\\alpha$ parameter. Also, the resolution is needed to evaluate the so called hemispheric helicity rule seen in the global $\\alpha$ parameter calculated from photospheric vector magnetograms \\citep{pcm94,pcm95,hagi04,nand06}. The availability of high resolution vector magnetograms from Hinode (SOT/SP), gives us the best opportunity so far to address this problem. The effect of polarimetric noise is expected to be negligible in the estimation of magnetic parameters \\citep{tiw09b} from these data. In this Letter we obtain an expression for the net current using a generalization of the expression obtained by \\cite{parker96}. We then proceed to measure this current from several vector magnetograms of nearly circular sunspots. We finally discuss the results and present our conclusions. ", "conclusions": "It is well known for astrophysical plasmas, that the plasma distorts the magnetic field and the curl of this distorted field produces a current by Ampere's law \\citep{parker79}. Parker's (1996) expectation of net zero current in a sunspot was chiefly motivated by the concept of a fibril structure for the sunspot field. However, he also did not rule out the possibility of vanishing net current for a monolithic field where the azimuthal component of the vector field in a cylindrical geometry declines faster than 1/$\\varpi$. While it is difficult to detect fibrils using the Zeeman effect notwithstanding the superior resolution of SOT on {\\it Hinode}, the stability and accuracy of the measurements have allowed us to detect the faster than 1/$\\varpi$ decline of the azimuthal component of the magnetic field, which in turn can be construed as evidence for the confinement of the sunspot field by the external plasma. The resulting pattern of curl {\\bf B} appears as a drop in net current at the sunspot boundary. If this lack of net current turns out to be a general feature of sunspot magnetic fields in the photosphere, then measurement of helicity from a global average of the force-free parameter becomes suspect. On the other hand, sunspots are evidently twisted at photospheric levels, as seen from the non-vanishing average twist angle as well as the SSA (Table 1). Although the existence of a global twist in the absence of a net current is possible for a monolithic sunspot field \\citep{baty00,arch04,fan04,aula05}, a fibril model of the sunspot field can accommodate a global twist even without a net current \\citep{parker96}. The spatial pattern of current density in a sunspot (e.g., left panel of Figure 4) is really a manifestation of the deformation of the magnetic field ($\\nabla\\times\\bf B$) by the forces applied by the plasma. The Lorentz force exerted by the field on the plasma produces an equal and opposite force by the plasma, thereby confining the field. Thus our analysis actually shows the pattern of the forces exerted by the plasma on the field. The sharp decline of the azimuthal field with radial distance thus shows the confinement of the sunspot magnetic field by the radial gradient of the plasma pressure. Theoretical understanding of the penumbral fine structure has improved considerably in recent times \\citep{thom02,weis04}. The onset of a convective instability for magnetic field inclination exceeding a critical value was proposed by \\cite{tilde03} and \\cite{hurl02}. A bifurcation in the onset \\citep{ruck95} could explain other features like hysteresis in the appearance of penumbra as a function of sunspot size. Numerical simulation of magneto-convection also steadily improved \\citep{hein07,remp09a}, culminating in very realistic production of penumbral field structure \\citep{remp09b}. It is possible, owing to the random and stochastic nature of convective structures, that no net twist in the simulated spot field would be produced by convection for negligible Coriolis force. If so, it would be very interesting to simulate magneto-convection in a twisted sunspot field. In this case, would the resulting fine structure mimic the observed ``curly interlocking combed'' structure of the penumbral magnetic field? If not, we must look elsewhere for explaining the ``curly interlocking combed'' structure. A twisted fibril bundle would then be a solution. Recent examples of filamentary penumbral structures based on such cluster models \\citep{sola93,spru06,scha06} have also been proposed. \\cite{parker96} also mentions the possibility of net currents in the corona, continuing down to the height where the first cleaving takes place. It would therefore be imperative to look for net currents at higher reaches of the solar atmosphere. This is very important because several theories of flares \\citep{melr95} and CME triggers \\citep{forb91,klie06} rely heavily on the existence of net currents in the corona above the sunspots. Future large ground based telescopes equipped with adaptive optics and multi spectral line capabilities would go a long way in addressing these issues. In the meantime, direct measurement of the global twist of sunspots using parameters like the SSA should serve as proxies for estimating the net currents of active regions in the corona. The SSA will also be a better parameter to base a fresh look at the hemispheric rule in photospheric chirality." }, "0910/0910.1081_arXiv.txt": { "abstract": "We report the first detection of the Zeeman effect in the 36 GHz Class I \\methanol\\ maser line. The observations were carried out with 13 antennas of the EVLA toward the high mass star forming region M8E. Based on our adopted Zeeman splitting factor of $z = 1.7$ Hz \\mG, we detect a line of sight magnetic field of $-31.3 \\pm 3.5$ mG and $20.2 \\pm 3.5$ mG to the northwest and southeast of the maser line peak respectively. This change in sign over a 1300 AU size scale may indicate that the masers are tracing two regions with different fields, or that the same field curves across the regions where the masers are being excited. The detected fields are not significantly different from the magnetic fields detected in the 6.7 GHz Class II \\methanol\\ maser line, indicating that \\methanol\\ masers may trace the large scale magnetic field, or that the magnetic field remains unchanged during the early evolution of star forming regions. Given what is known about the densities at which 36 GHz \\methanol\\ masers are excited, we find that the magnetic field is dynamically significant in the star forming region. ", "introduction": "\\label{sINTRO} Magnetic fields likely play an important role in the star formation process (e.g., \\citealt{tt2008}, and references therein). The nature of this role is not yet clear, however, primarily due to the scarcity of observational data on magnetic fields in star forming regions. The Zeeman effect remains the most direct method for measuring magnetic field strengths (e.g., \\citealt{rmc99}). Observations of the Zeeman effect in H I and OH thermal lines have revealed the strength of the magnetic field in the lower density envelopes of molecular clouds (e.g., \\citealt{bt01}). Observations of the Zeeman effect in \\water\\ masers, on the other hand, offer a window into magnetic fields in the highest density regions ($n \\sim 10^9~\\cm{-3}$) of star forming regions (\\citealt{sarma2008}; \\citealt{vdv06}). Interstellar masers, being compact and intense, are extremely effective probes of young star forming regions. Recently, \\citet{wv2008} carried out the first systematic study of the Zeeman effect in the 6.7 GHz methanol maser line. Such (6.7 GHz) masers are examples of Class II methanol masers, which are known to be probes of the early phases of high mass star forming regions. Yet another class of masers, namely Class I methanol masers, may probe even earlier phases of star forming regions (\\citealt{pp2008}). This appears to be the case for M8E, which is a high mass star forming region located at $l = 6\\arcdeg.05$, $b = -1\\arcdeg.45$, at a distance of 1.5 kpc (\\citealt{scf1984}; \\citealt{gg76}). M8E is comprised of a bright infrared source (M8E-IR), and a radio \\ion{H}{2}\\ region about 8$\\arcsec$ to its northwest (\\citealt{scf1984}; \\citealt{dw2009}). Low spatial resolution CO observations by \\citet{wright77} showed that M8E is embedded in a dense molecular core. At higher resolution, \\citet{mhs92} mapped a bipolar outflow in the $^{12}$CO emission line (\\vLSR = 11 \\kmS), with blue and red wings spread over 50 \\kmS. Based on their VLTI observations, \\citet{linz2009} concluded that M8E-IR contains a central protostar of mass 10-15 \\Msun. In this paper, we report the \\textsl{first} detection of the Zeeman effect in the 36 GHz Class I \\methanol\\ maser line. The observations were carried out with 13 antennas of the Expanded Very Large Array (EVLA) toward the high mass star-forming region M8E, and represent an early success of the EVLA. In \\S\\ \\ref{sODR}, we present details of the observations and reduction of the data. The analysis involved in extracting magnetic field information from the Zeeman effect is given in \\S\\ \\ref{sANAL}. The results are presented and discussed in \\S\\ \\ref{sR}. ", "conclusions": "\\label{sCONC} We have detected for the first time the Zeeman effect in the 36 GHz Class I \\methanol\\ maser line toward the high mass star forming region M8E. The detected values for the line-of-sight magnetic field are $-31.3 \\pm 3.5$ mG and $20.2 \\pm 3.5$ mG. There may be systematic bias in these values due to the assumed Zeeman splitting factor of $z = 1.7$ Hz \\mG, which is based on laboratory experiments on the 25 GHz \\methanol\\ maser line. Our detected values are similar to the magnetic fields detected in 6.7 GHz Class II \\methanol\\ masers by \\citet{wv2008}. This suggests that \\methanol\\ masers may trace the large-scale magnetic field; alternatively, the magnetic field may remain the same during the early evolution of the star forming region between the time when Class I and Class II masers are excited. Our detected values for \\Blos\\ imply that the magnetic field is dynamically significant to these regions. Our detection gives rise to a host of questions that need further observational and theoretical work. We are planning future observations of the Zeeman effect in more 36 GHz masers and at higher resolution. We are also planning for observations of 44 GHz masers with the aim of detecting the Zeeman effect. As mentioned above, the 44 GHz line constitutes the other prominent Class I \\methanol\\ maser line. We are also planning to carry out high angular resolution observations of the Zeeman effect in regions containing both Class I and Class II masers, in order to compare the strength of magnetic fields between the regions traced by these classes. From theorists, we seek detailed models for Class I masers in order to understand their properties better, especially the range of densities at which they are excited. Also, there is a pressing need for laboratory studies of the Zeeman splitting coefficient for \\methanol\\ masers. The resolution of some or all of these issues will open up yet another important window into the early stages of high mass star forming regions." }, "0910/0910.4989_arXiv.txt": { "abstract": "We present axisymmetric hydrodynamical simulations of the long-term accretion of a rotating gamma-ray burst progenitor star, a ``collapsar,'' onto the central compact object, which we take to be a black hole. The simulations were carried out with the adaptive mesh refinement code FLASH in two spatial dimensions and with an explicit shear viscosity. The evolution of the central accretion rate exhibits phases reminiscent of the long GRB $\\gamma$-ray and X-ray light curve, which lends support to the proposal by \\citet{Kumar:08a,Kumar:08b} that the luminosity is modulated by the central accretion rate. In the first ``prompt'' phase characterized by an approximately constant accretion rate, the black hole acquires most of its final mass through supersonic quasiradial accretion occurring at a steady rate of $\\sim 0.2~M_\\odot ~\\textrm{s}^{-1}$. After a few tens of seconds, an accretion shock sweeps outward through the star. The formation and outward expansion of the accretion shock is accompanied with a sudden and rapid power-law decline in the central accretion rate $\\dot M\\propto t^{-2.8}$, which resembles the $L_{\\rm X}\\propto t^{-3}$ decline observed in the X-ray light curves. The collapsed, shock-heated stellar envelope settles into a thick, low-mass equatorial disk embedded within a massive, pressure-supported atmosphere, similar to the picture proposed by \\citet{Begelman:08} for ``quasistars.'' After a few hundred seconds, the inflow of low-angular-momentum material in the axial funnel reverses into an outflow from the surface of the thick disk. Meanwhile, the rapid decline of the accretion rate slows down, or even settles a in steady state with $\\dot M\\sim 5\\times10^{-5}~M_\\odot~\\textrm{s}^{-1}$, which resembles the ``plateau'' phase in the X-ray light curve. While the duration of the ``prompt'' phase depends on the resolution in our simulations, we provide an analytical model taking into account neutrino losses that estimates the duration to be $\\sim 20~\\textrm{s}$. The model suggests that the steep decline in GRB X-ray light curves is triggered by the circularization of the infalling stellar envelope at radii where the virial temperature is below $10^{10} ~\\textrm{K}$, such that neutrino cooling shuts off and an outward expansion of the accretion shock becomes imminent; GRBs with longer prompt $\\gamma$-ray emission have more slowly rotating envelopes. ", "introduction": "\\label{sec:intro} \\setcounter{footnote}{0} Observations of long gamma-ray bursts (GRBs) carried out with the NASA {\\it Swift} satellite have shown that the $\\gamma$-ray prompt emission ceases after about a minute in the observer frame, corresponding to tens of seconds in the rest frame of the progenitor star. The $\\gamma$-ray light curve, converted to a fiducial X-ray spectral band, smoothly joins the X-ray light curve, which declines rapidly, ($t^{-3}$ or faster) lasting for about $80$ to $300~\\textrm{s}$ \\citep{Tagliaferri:05,Nousek:06,OBrien:06}. The rapid decline is often followed by a phase, from about $\\sim 10^3$ to $10^4~\\textrm{s}$, during which the X-ray flux is roughly constant or declines more slowly with time. The X-ray light curves of some GRBs exhibit ``flares'' where the flux increases suddenly by a factor of $\\lesssim 10^2$ and drops precipitously, with the rise and decline associated with the flare occurring on a time scale much shorter than the age of the burst \\citep[see, e.g.,][]{Burrows:05,Falcone:06}. Following about $\\sim10^3-10^4~\\textrm{s}$, a more rapid decline of the luminosity resumes \\citep[see, e.g.,][and references therein]{ZhangB:06}, and occasionally steepens further at $\\sim10^4-10^5~\\textrm{s}$ \\citep[e.g.,][]{Vaughan:06}. The goal of the present work is to utilize two-dimensional hydrodynamic simulations to test the hypothesis \\citep{Kumar:08a,Kumar:08b} that this characteristic structure of the X-ray light curve, which was summarized by \\citet{ZhangB:06}, reflects a modulation in the rate of central accretion of a rotating progenitor star onto a black hole or a neutron star, as in the collapsar model of GRBs \\citep{Woosley:93,MacFadyen:99,MacFadyen:01,Woosley:06b}. We do not attempt to explore the implications of the potential presence of a magnetosphere, as in the magnetar model for GRBs \\citep[e.g.,][]{Duncan:92,Wheeler:00,ZhangB:01,Thompson:04,Komissarov:07,Bucciantini:07,Bucciantini:09}. We will attempt to gain insight in the origin of the steady $\\gamma$-ray luminosity (the prompt phase which we will refer to as ``Phase 0''), the rapid decline in the X-ray light curve (Phase I in the nomenclature of \\citealt{ZhangB:06}), and phase of quasi-steady luminosity or slow decline (Phase II). We will briefly attempt to extrapolate the results of our simulations to the subsequent steeper decline phases (Phases III and IV). \\citet{Kumar:08a,Kumar:08b} obtained the key features of the $\\gamma$-ray and X-ray light curve by estimating central accretion rate resulting from the free (i.e., ballistic) infall of a rotating progenitor star. In this picture, the material that has sufficient initial angular momentum to circularize outside of the innermost stable circular orbit (ISCO) of the black hole, forms a disk in the equatorial plane, and subsequently accretes via disk accretion \\citep{Narayan:01}. If the luminosity is then assumed to be proportional to the central accretion rate, and if the distance of the $\\gamma$-ray or X-ray emitting region from the center of the star is assumed to be approximately independent of time on time scales $10-10^5~\\textrm{s}$, an accretion model directly translates into a synthetic light curve that can be compared with an observed light curve. \\citet{Kumar:08a,Kumar:08b} have shown that with the simplest accretion model involving ballistic infall onto the midplane (assumption also made by \\citealt{Janiuk:08} and \\citealt{Cannizzo:09}) and subsequent disk accretion, the mapping of the mass accretion history onto the light curve provides a powerful insight into the stratification and angular momentum structure of the progenitor star. In their ballistic infall model, \\citet{Kumar:08a,Kumar:08b} were not able to discriminate between models in which the quasi-steady activity in Phase II arose from disk accretion, or from late-time accretion from an extended stellar envelope. Departures from ballistic infall are expected if the infalling material passes through an accretion shock \\citep[see, e.g.,][]{MacFadyen:99,Lee:06,Nagataki:07,LopezCamara:09}, or if the disk launches a thermal \\citep[][]{MacFadyen:03a,MacFadyen:03b,Kohri:05} or magnetohydrodynamic \\citep[e.g.,][]{Proga:03c} outflow (``wind'') that can interfere with the infall. The existence of the outflow is particularly interesting because of the potential for nucleosynthesis in the free neutron-rich outflow launched from the inner part of the disk \\citep[see, e.g.,][]{Pruet:03,Surman:06,Fujimoto:07,Nagataki:07,Maeda:09} and because of the potential that the outflow can deplete the accreting stellar envelope and limit the envelope mass that is accreted onto the central black hole. During the first $\\sim10^2~\\textrm{s}$ following the formation of the central black hole when the accretion rate is $\\dot M\\gg 10^{-3}M_\\odot~\\textrm{s}^{-1}$ (the precise condition depends on the black hole spin and shear stress-to-pressure ratio $\\alpha$ in the disk), the inner accretion disk cools by neutrino emission and nuclear photodisintegration and accretes in a radiatively efficient fashion, except for in the very inner, optically thick region \\citep[e.g.,][]{Popham:99,Narayan:01,DiMatteo:02,Chen:07}. Instabilities in the thin disk have been cited as a candidate class of mechanisms that could produce the observed X-ray flares \\citep{Perna:06,Lazzati:07,Lazzati:08} and could also produce detectable gravitational radiation \\citep{Piro:07}. Our global axisymmetric models are the necessary stepping stone toward the substantially more computationally demanding three-dimensional simulations that will be required to pin down any nonaxisymmetric instabilities in the accreting collapsar \\citep[see, e.g.,][]{Rockefeller:06}. We employ two-dimensional unmagnetized hydrodynamic simulations of the collapse, circularization, and accretion of a stellar envelope onto a central point mass, which we assume to be a black hole; relativistic corrections to the gravitational potential are ignored in our simulations since the innermost grid point lies at over $20$ Schwarzschild radii in the simulation extending to the smallest radius from the black hole. The torque and dissipation arising from the $R-\\phi$ component of the magnetic stress is emulated with a Navier-Stokes term parameterized by an $\\alpha$-viscosity prescription. For comparison with the X-ray light curve, we measure the central accretion rate. We track the flow of mass and energy at spherical radii $10^8~\\textrm{cm}\\lesssim r\\lesssim 10^{11}~\\textrm{cm}$ and interpret the results in view of the existing knowledge on radiatively-inefficient accretion flows. We observe an outflow and measure the rate at which the accreting stellar envelope is lost to the outflow. The mechanics of post-core-collapse accretion and outflows is key to estimating the final mass of the black hole and the nucleosynthetic composition of the ejected matter \\citep[e.g.,][and references therein]{Zhang:08}. The method that we develop here can in future be utilized to estimate the masses of the black holes resulting from the collapse of massive, initially metal-poor ``Population III'' stars as well as from the collapse of the even more massive, hypothetical ``supermassive stars,'' in the presence of rotation. In this work we do not simulate the neutrino-cooled disk, and in the simulations simply impose that the mass that crosses the innermost cylindrical radius of our simulation, $R_{\\rm min} = (0.5-2) \\times 10^{8} ~\\textrm{cm}$, is instantaneously incorporated inside the black hole and does not provide any further energetic feedback while at radii $R5\\times10^7~\\textrm{cm}$ and occurs at quasi-constant rate of $\\sim0.2~M_\\odot~\\textrm{s}^{-1}$. The end of this phase is marked by the formation of an accretion shock at the smallest resolved radii. The shock immediately propagates radially outwards through the supersonically infalling stellar envelope. Simultaneously with the formation and the outer movement of the accretion shock, the accretion rate drops suddenly and precipitously. We argue that the somewhat late onset of the accretion shock is an artifact of our not resolving the innermost two decades in radius outside the black hole's gravitational radius. We supplement the simulations with an analytical model of the innermost accretion disk not resolved in the simulations, and suggest that the accretion shock forms early, within a fraction of the first second of the formation of the black hole, as several published simulations of the innermost neutrino-cooled region have shown, but only starts to propagate outward after $20~\\textrm{s}$, when the material that is reaching the equatorial plane has enough angular momentum to circularize at radii where the virial temperature is below $\\sim 10^{10}~\\textrm{K}$ and the cooling by neutrino emission is suppressed. During the second phase characterized by a steep decline $\\propto t^{-2.7}$ of the accretion rate that lasts $\\sim 500~\\textrm{s}$, the accretion shock sweeps through the star, but a supersonic accretion of the shocked fluid in the axial funnel region proceeds unabated, at least in our simulations where the funnel has not been heated to high temperatures by the relativistic jet. The thick disk containing rotationally-supported and pressure-supported fluid is convective; a high-entropy outflow from the inner, rotationally-supported region follows the accretion shock on its traversal through the star but remains bound within the star and appears to form a large-scale circulation pattern. The steepness of the accretion rate decline seems to be the consequence of a rapid hydrodynamic readjustment of the shocked, convective, and circulating stellar envelope. The steep decline of the accretion rate slows down or stalls after $\\sim 600~\\textrm{s}$, which appears to reflect the settling of a fraction of the stellar envelope in the state of near-hydrostatic equilibrium. The inner, rotationally supported thick disk contains $\\sim 1\\%$ of the mass of the unaccreted envelope and extends to $\\sim 3\\times10^9~\\textrm{cm}$. The thick disk is surrounded by a much more massive, \\emph{pressure-supported} atmosphere, which acts as a mass supply to the thick disk. At no point do we find evidence for the extended thin disk envisioned by \\citet{Cannizzo:09}. The fluid above and below the thick disk is mostly unbound and the simulations thus exhibit a form of a ``disk wind.'' We speculate that depletion of the envelope through accretion onto the black hole or mass loss in thermal outflows or winds could be responsible for the renewed steepening of the GRB X-ray light curve after $10^3-10^4~\\textrm{s}$. More speculatively, the additional steepening of the light curve occasionally observed after $10^4-10^5~\\textrm{s}$ could be due to a pervasive thermal or radiatively-driven mass loss in the outer layers of the atmosphere." }, "0910/0910.4171_arXiv.txt": { "abstract": "We present Chandra snapshot observations of the first large X-ray sample of optically identified fossil groups. For 9 of 14 candidate groups, we are able to determine the X-ray luminosity and temperature, which span a range typical of large ellipticals to rich groups of galaxies. We discuss these initial results in the context of group IGM and central galaxy ISM evolution, and we also describe plans for a deep X-ray follow-up program. ", "introduction": " ", "conclusions": "" }, "0910/0910.4582_arXiv.txt": { "abstract": "The centers of ellipticals and bulges are formed dissipationally, via gas inflows over short timescales -- the ``starburst'' mode of star formation. Recent work has shown that the surface brightness profiles, kinematics, and stellar populations of spheroids can be used to separate the dissipational component from the dissipationless ``envelope'' made up of stars formed over more extended histories in separate objects, and violently assembled in mergers. Given high-resolution, detailed observations of these ``burst relic'' components of ellipticals (specifically their stellar mass surface density profiles), together with the simple assumptions that some form of the Kennicutt-Schmidt law holds and that the burst was indeed a dissipational, gas-rich event, we show that it is possible to invert the observed profiles and obtain the time and space-dependent star formation history of each burst. We perform this exercise using a large sample of well-studied spheroids, which have also been used to calibrate estimates of the ``burst relic'' populations. We show that the implied bursts scale in magnitude, mass, and peak star formation rate with galaxy mass in a simple manner, and provide fits for these correlations. The typical burst mass $M_{\\rm burst}$ is $\\sim10\\%$ of the total spheroid mass; the characteristic starburst timescale implied is a nearly galaxy-mass independent $t_{\\rm burst}\\sim10^{8}\\,$yr; the peak SFR of the burst is $\\sim M_{\\rm burst}/t_{\\rm burst}$; and bursts decay subsequently in power-law fashion as $\\dot{M}_{\\ast}\\propto t^{-2.4}$. As a function of time, we obtain the spatial size of the starburst; burst sizes at peak activity scale with burst mass in a manner similar to the observed spheroid size-mass relation, but are smaller than the full galaxy size by a factor $\\sim10$; the size grows in time as the central, most dense regions are more quickly depleted by star formation as $R_{\\rm burst}\\propto t^{0.5}$. Combined with observational measurements of the nuclear stellar population ages of these systems -- i.e.\\ the distribution of times when these bursts occurred -- it is possible to re-construct the dissipational burst contribution to the distribution of SFRs and IR luminosity functions and luminosity density of the Universe. We do so, and show that these burst luminosity functions agree well with the observed IR LFs at the brightest luminosities, at redshifts $z\\sim0-2$. At low luminosities, however, bursts are always unimportant; the transition luminosity between these regimes increases with redshift from the ULIRG threshold at $z\\sim0$ to Hyper-LIRG thresholds at $z\\sim2$. At the highest redshifts $z\\gtrsim2$, we can set strict upper limits on starburst magnitudes, based on the maximum stellar mass remaining at high densities at $z=0$, and find tension between these and estimated number counts of sub-millimeter galaxies, implying that some change in bolometric corrections, the number counts themselves, or the stellar IMF may be necessary. At all redshifts, bursts are a small fraction of the total SFR or luminosity density, approximately $\\sim5-10\\%$, in good agreement with estimates of the contribution of merger-induced star formation. ", "introduction": "\\label{sec:intro} Understanding the global star-formation history of the Universe remains an important unresolved goal in cosmology. Of particular interest is the role played by mergers in driving star formation and/or the infrared luminosities of massive systems. A wide range of observations support the view that violent, dissipational events (e.g.\\ gas-rich mergers) are important to galaxy evolution, and in particular that the central, dense portions of galaxy bulges and spheroids must be formed in such events; but less clear is their contribution to the global star formation process. In the local Universe, the population of star-forming galaxies appears to transition from ``quiescent'' (undisturbed) disks -- which dominate the {\\em total} star formation rate/IR luminosity density -- at the luminous infrared galaxy (LIRG) threshold $10^{11}\\,\\lsun$ ($\\dot{M}_{\\ast}\\sim 10-20\\,\\msun\\,{\\rm yr^{-1}}$) to systems that are clearly merging and violently disturbed at a few times this luminosity. The most intense starbursts at $z=0$, ultraluminous infrared galaxies (ULIRGs; $L_{\\rm IR}>10^{12}\\,\\lsun$), are invariably associated with mergers \\citep[e.g.][]{joseph85,sanders96:ulirgs.mergers}, with dense gas in their centers providing material to feed black hole growth and to boost the concentration and central phase space density of merging spirals to match those of ellipticals \\citep{hernquist:phasespace,robertson:fp}. Various studies have shown that the mass involved in these starburst events is critical to explain the relations between spirals, mergers, and ellipticals, and has a dramatic impact on the properties of merger remnants \\citep[e.g.,][]{LakeDressler86,Doyon94,ShierFischer98,James99, Genzel01,tacconi:ulirgs.sb.profiles,dasyra:mass.ratio.conditions,dasyra:pg.qso.dynamics, rj:profiles,rothberg.joseph:kinematics,hopkins:cusps.ell,hopkins:cores}. At high redshifts, bright systems dominate more and more of the IR luminosity function \\citep[e.g.][]{lefloch:ir.lfs,perezgonzalez:ir.lfs,caputi:ir.lfs,magnelli:z1.ir.lfs}. Merger rates increase rapidly \\citep[by a factor $\\sim10$ from $z=0-2$; see e.g.][and references therein]{hopkins:merger.rates}, leading to speculation that the merger rate evolution may in fact drive the observed evolution in the cosmic SFR density, which also rises sharply in this interval \\citep[e.g.][and references therein]{hopkinsbeacom:sfh}. However, many LIRGs at $z\\sim1$, and possibly ULIRGs at $z\\sim2$, appear to be ``normal'' galaxies, without dramatic morphological disturbances associated with the local starburst population or large apparent AGN contributions \\citep{yan:z2.sf.seds,sajina:pah.qso.vs.sf, dey:2008.dog.population,melbourne:2008.dog.morph.smooth, dasyra:highz.ulirg.imaging.not.major}. At the same time, even more luminous systems appear, including large numbers of Hyper-LIRG (HyLIRG; $L_{\\rm IR}>10^{13}\\,\\lsun$) and bright sub-millimeter galaxies \\citep[e.g.][]{chapman:submm.lfs,younger:highz.smgs,casey:highz.ulirg.pops}. These systems exhibit many of the traits more commonly associated with merger-driven starbursts, including morphological disturbances, and may be linked to the emergence of massive, quenched (non star-forming), compact ellipticals at times as early as $z\\sim2-4$ \\citep{papovich:highz.sb.gal.timescales, younger:smg.sizes,tacconi:smg.maximal.sb.sizes, schinnerer:submm.merger.w.compact.mol.gas, chapman:submm.halo.clustering,tacconi:smg.mgr.lifetime.to.quiescent}. In a series of papers, \\citet{hopkins:cusps.mergers, hopkins:cusps.ell,hopkins:cores} (hereafter H08c, H09b, H09e, respectively), the authors combined simulation libraries of galaxy mergers with observations of nearby ellipticals to develop a methodology to empirically separate spheroids into their two dominant physical components. First, a dissipationless component -- i.e.\\ an ``envelope,'' formed from the violent relaxation/scattering of stars already present in merging stellar disks that contribute to the remnant. Because disks are extended, with low phase-space density (and collisionless processes cannot raise this phase-space density), these stars will necessarily dominate the profile at large radii, hence the envelope, with a low central density. Second, a dissipational or ``burst'' component -- i.e.\\ a dense stellar relic formed from gas which lost its angular momentum and fell into the nucleus of the remnant, turning into stars in a compact central starburst like those in e.g.\\ local ULIRGs (although these represent the most extreme cases). Because gas radiates, it can collapse to very high densities, and stars formed in a starburst will dominate the profile within radii $\\sim0.5-1\\,$kpc, accounting for the high central densities of ellipticals. The gas will reflect that brought in from merging disks, but could in principle also be augmented by additional cooling or stellar mass loss in the elliptical \\citep[see e.g.][]{ciottiostriker:recycling}. In subsequent mergers, these two stellar components will act dissipationlessly, but the segregation between the two is sufficient that they remain distinct even after multiple, major ``dry'' re-mergers: i.e.\\ one can still, in principle, distinguish the dense central stellar component that is the remnant of the combined dissipational starburst(s) from the less dense outer envelope that is the remnant from low-density disk stars. In H08c, the methodology for empirically separating these components in observed systems is presented, and tested this on samples of nearby merger remnants from \\citet{rj:profiles}. Comparison with other, independent constraints such as stellar population synthesis models \\citep{titus:ssp.decomp, reichardt:ssp.decomp,michard:ssp.decomp} and galaxy abundance profiles \\citep{foster:metallicity.gradients.prep, mcdermid:sauron.profiles,sanchezblazquez:ssp.gradients}, as well as direct comparison of simulations with observed surface brightness profiles, galaxy shapes, and kinematics are used to demonstrate that the approach can reliably extract the dissipational component of the galaxy (see \\S~\\ref{sec:methods:obs}). If this general scenario is correct, then the empirically identified burst relic components in local spheroids represent a novel and powerful new constraint on the history and nature of dissipational starbursts. In this paper, we present and develop these constraints -- and in particular, we show that they are not merely integral constraints on the masses and sizes of bursts. We show that the unique nature of dissipational star formation, together with the existence of some Kennicutt-Schmidt type relation between gas surface densities and star formation rates, means that the relic profiles can be {\\em inverted} to obtain the full time and radius-dependent star formation history of each galaxy starburst (\\S~\\ref{sec:methods}). We present the resulting, derived burst star formation histories for samples of hundreds of local, well-observed galaxy spheroids. We show that such starbursts follow broadly similar time-dependent behavior, with characteristic starburst star formation rates, durations, rise and decay rates, and spatial sizes that scale with galaxy mass and other properties according to simple scaling laws across $\\sim5$ decades in starburst and spheroid mass (\\S~\\ref{sec:results:properties}). Combining these inferred star formation histories (and resulting burst lightcurves) with empirical determinations of the ages of each starburst, we can reconstruct the starburst history of the Universe, including e.g.\\ the luminosity functions of such dissipational bursts at all redshifts, and their contribution to the global star formation rate density (\\S~\\ref{sec:results:history}). We discuss the uncertainties in this approach, compare with the previous, independent attempts to constrain these quantities via other observational methods, and summarize our results in \\S~\\ref{sec:discussion}. Throughout, we adopt a $\\Omega_{\\rm M}=0.3$, $\\Omega_{\\Lambda}=0.7$, $h=0.7$ cosmology and a \\citet{chabrier:imf} stellar IMF, but these choices do not affect our conclusions. ", "conclusions": "\\label{sec:discussion} It has long been believed that the centers of galaxy spheroids must be formed in dissipational starburst events, such that gas in the outer parts of a galaxy must lose angular momentum rapidly, and fall in on roughly a dynamical time to the galaxy center. Recently, observations have shown that it is possible to robustly separate the ``burst'' component of galaxy profiles from the outer, violently relaxed component owing to the scattering of progenitor galaxy stars formed over earlier, more extended periods. As detailed, high-resolution observations of e.g.\\ spheroid stellar mass profiles, shapes, kinematics, and stellar population properties improve, such decompositions become increasingly robust and applicable to a wide range of systems. In this paper, we use these observations as a novel, independent constraint on the nature of galactic starbursts. We stress that by ``starburst'' here, we do {\\em not} mean simply any high-SFR system; rather, we refer specifically to star formation in central, usually sub-kpc scale bursts owing to large central gas concentrations driven by angular momentum loss as described above. These are commonly associated with galaxy-galaxy mergers, which have long been known to efficiently drive starbursts and violent relaxation \\citep[e.g.][]{lynden-bell67, toomre77,barnes.hernquist.91,barneshernquist96,barnes:review, mihos:starbursts.94,mihos:starbursts.96}. But they could also owe to any other sufficiently violent process; e.g.\\ gas inflows owing to disk bars or during dissipational collapse. Regardless of origin, we can robustly identify the starburst relics and stellar mass surface density profiles in well-observed spheroids at $z=0$. This methodology and tests of its accuracy have been discussed in \\citet{hopkins:cusps.mergers,hopkins:cusps.ell,hopkins:cores,hopkins:cusps.fp}. This alone is a powerful integral constraint on the star formation histories of these systems. But we show that they can in fact be used to provide more detailed information, by allowing for the inversion and recovery of the star formation history of each burst. We can thus use local observations to recover the full time-dependent and scale-dependent star formation history of each burst, if we couple these observed constraints to simple, well-motivated assumptions. Namely, that some form of the Kennicutt-Schmidt relation (between star formation rate and gas surface density) applied in the formation of these stars, and that they formed in dissipational (i.e.\\ rapid and initially gas-rich, on these scales) events. Both assumptions are directly motivated by observations, but we also consider how uncertainties in them translate into uncertainties in the resulting constraints. Performing this exercise, we recover a large number of empirically determined parameters of starbursts, and quantify how they scale as a function of mass and other properties. This includes the total mass of the starburst and the time-dependent SFR. In general, we find that the constrained starbursts can be reasonably well-approximated by a simple power-law behavior (Equation~\\ref{eqn:mdot.time}), rising and decaying from a characteristic maximum SFR with characteristic half-life $t_{\\rm burst}$ and a typical late-time power-law slope of $\\dot{M}_{\\ast}\\propto (t/t_{\\rm burst})^{-\\beta}$ where $\\beta \\sim 2.0-3.0$. The characteristic burst mass $M_{\\rm burst}$ is $\\sim10\\%$ of the total spheroid mass, but it scales weakly with galaxy mass in a manner similar to how disk galaxy gas fractions scale with their stellar masses (expected if disks are the pre-merger progenitors of spheroids), $M_{\\rm burst}\\approx 1/(1+[M_{\\ast}/10^{9.15}\\,\\msun]^{0.4})$. Detailed discussion of this trend, and its consequences for the global structure and kinematics of spheroids, are presented in \\citet{hopkins:cusps.fp}. However, it already makes it clear that bursts should not dominate the SFR density. They are a small fraction $\\sim10\\%$ of all stars in spheroids (let alone all stars). That does not mean that bursts are unimportant, however. It is clear that they control many of the properties of galaxies \\citep{cox:kinematics,robertson:fp, naab:gas,onorbe:diss.fp.details,ciotti:dry.vs.wet.mergers, jesseit:kinematics,jesseit:merger.rem.spin.vs.gas, covington:diss.size.expectation, hopkins:cusps.ell,hopkins:cores}, and they can account for the short-lived, highest-SFR systems in the Universe. Moreover, the scatter in burst mass is significant, $\\sim0.3-0.4\\,$dex, which is critical for explaining the most extreme starbursts in the Universe, which require both large absolute galaxy masses and gas fractions to reach the very high burst masses required. The typical starburst timescale implied from the combination of observed surface densities and the \\citet{kennicutt98} law is a nearly galaxy-mass independent $t_{\\rm burst}\\sim10^{8}\\,$yr. Both the value of this timescale, and the weak scaling with galaxy and/or burst mass, agree well with the dynamical times in the central $\\sim$kpc of galaxies. As above, though, there is considerable scatter of order $\\sim0.2\\,$dex. Given such a short starburst timescale, and the relatively small total mass fractions involved in starbursts, it naturally follows that starbursts will represent only a small fraction of the star-forming galaxies at any stellar mass, at a particular instant. Given that the average number of bursts per galaxy is not large, the duty cycle of bursts should be $\\sim t_{\\rm burst}/t_{\\rm Hubble}$, or $\\sim1-5\\%$ from $z=0-3$. Indeed, observations have shown that most galaxies at these redshifts lie on a normal star-forming sequence, without a large $\\sim1\\,\\sigma$ scatter from e.g.\\ merger-induced bursts \\citep{noeske:2007.sfh.part1, noeske:sfh,papovich:ssfr,bell:morphology.vs.sfr}. The small duty cycle here means that even if the burst SFR enhancement is large, it will not violate these constraints (appearing only at the $\\sim2-3\\,\\sigma$ level in the wings of the SFR distribution at a given mass). These conclusions are supported by independent evidence from observational stellar population synthesis studies. Specific comparisons to the objects considered here, where available, are presented in detail in \\citet{hopkins:cusps.ell,hopkins:cores} and \\citep{foster:metallicity.gradients.prep}. It is well-established that constraints from abundances require the central portions of spheroids be formed in a similar, short timescale. And detailed decomposition of stellar populations into burst plus older stellar populations have yielded consistent results for the typical burst fractions and sizes \\citep{titus:ssp.decomp,schweizer:7252, reichardt:ssp.decomp,michard:ssp.decomp}. As should be expected from the generic behavior above, bursts peak at SFRs of $\\sim M_{\\rm burst}/t_{\\rm burst}$, which follows $M_{\\rm burst}$ (and hence total spheroid mass) in a close-to-linear relation. The most massive local ellipticals -- especially those with total stellar masses of $\\gtrsim10^{12}\\,\\msun$ -- had extreme peak SFRs of $>1000\\,\\msun\\,{\\rm yr^{-1}}$. Thus, it is at least possible that some local systems reached the highest SFRs inferred for massive, high-redshift starburst galaxies \\citep{papovich:highz.sb.gal.timescales, chapman:submm.lfs,walter:2009.hyperlirg.in.highz.qso.host}. More moderate, but still massive ellipticals with $M_{\\ast}\\sim 1-5\\times10^{11}\\,\\msun$, reached a range of peak SFRs from $\\sim 30-500\\,\\msun\\,{\\rm yr^{-1}}$, corresponding to their forming fractions from $\\sim5-20\\%$ of their total masses in starbursts. These match well with the observed SFRs in more typical, local and $z\\sim1$ starbursts (in $\\sim L_{\\ast}$ galaxies), which have been specifically associated with mergers driving gas to galaxy centers, and forming the appropriate nuclear mass concentrations to explain e.g.\\ elliptical kinematics, sizes, phase space densities, and fundamental plane scalings \\citep{cox:kinematics, naab:gas,robertson:fp,jesseit:merger.rem.spin.vs.gas,hopkins:cusps.mergers, hopkins:cusps.fp,hopkins:cusps.evol}. We similarly quantify starburst spatial sizes as a function of their mass, peak SFR, and time. Starburst sizes scale with starburst mass in a similar fashion as the spheroid mass-size relation, but are smaller than their host spheroids by a fraction similar to their mass fraction. The most massive starbursts reach half-SFR (i.e.\\ half-light, in IR or mm wavelengths) size scales of $\\sim 1-5\\,$kpc. For typical spatial distributions, this implies total spatial extents of strong emission up to $\\sim10\\,$kpc. These size scales also agree well with observations of the most massive high-redshift starbursting systems \\citep{younger:smg.sizes, tacconi:smg.maximal.sb.sizes, schinnerer:submm.merger.w.compact.mol.gas}, and of massive, compact ellipticals formed at high redshift, believed to be the relics of such starbursts \\citep[with, at that time, little envelope of dissipationless, low-density material yet accreted;][]{vandokkum:z2.sizes, cimatti:highz.compact.gal.smgs,trujillo:compact.most.massive, bezanson:massive.gal.cores.evol,hopkins:r.z.evol}. For more typical, $\\sim L_{\\ast}$ starbursts, sizes range from $\\sim0.1-1\\,$kpc, also similar to those observed in merging systems \\citep{scoville86,sanders:review, hibbard.yun:excess.light,tacconi:ulirgs.sb.profiles,laine:toomre.sequence, rj:profiles}. The size difference between these and the most extreme objects follows from the much larger gas supply involved -- there is no dramatic difference in the relic starburst mass distribution {\\em shapes}. Some claims have been made that the large sizes of high-redshift starbursts could imply that they are not scaled up analogues of local extreme starbursts; but we find here that they correspond naturally. Scaling up a starbursting system in the starburst mass fraction will not preserve spatial size, but rather will scale along the starburst size-mass relation here, which appears to be smooth and continuous from the smallest starbursts with masses $\\sim10^{7}\\,\\msun$ to the largest with masses $>10^{11}\\,\\msun$. The origin of the size-mass scaling is of considerable physical interest. It has been proposed that in such starbursts the Eddington limit from radiation pressure on dusty gas sets a universal maximum central surface density, over all mass scales, from which this follows \\citep{hopkins:maximum.surface.densities}. The important constraint, from our analysis, is that there is no discontinuity, and we provide the scaling relations that any such model must satisfy. Combining these constraints with observational measurements of the nuclear stellar population ages of these systems -- i.e.\\ the distribution of times when these bursts occurred -- we show that it is possible to re-construct the dissipational burst contribution to the distribution of SFRs and IR luminosity functions and luminosity density of the Universe. We show that the burst luminosity functions agree well with the observed IR LFs at the brightest luminosities, at redshifts $z\\sim0-2$. At low luminosities, however, bursts are always unimportant, as expected from their short duty cycles, noted above. Although the burst luminosity functions rise with redshift, they always represent low space densities, and the overall LF evolves rapidly. As such, the transition luminosity above which bursts dominate the IR LFs and SFR distributions increases with redshift from the ULIRG threshold at $z\\sim0$ to Hyper-LIRG thresholds at $z\\sim2$. This appears to agree well with recent estimates of the transition between normal star formation and mergers, along the observed luminosity functions. Systematic morphological studies at low redshifts \\citep{sanders96:ulirgs.mergers} yield the conventional wisdom that -- locally -- the brightest LIRGs and essentially all ULIRGs are merging systems (see also references in \\S~\\ref{sec:intro}). At high redshifts, similar studies have now been performed \\citep[see e.g.][and references therein]{tacconi:smg.mgr.lifetime.to.quiescent}. They too find that the brightest sources are almost exclusively mergers, but with a transition point an order-of-magnitude higher in luminosity. Other morphological studies at intermediate redshifts $z\\sim0.4-1.4$ have reached similar conclusions \\citep{bridge:merger.fractions}. Integrating these luminosity functions, we find the burst contribution to the SFR and IR luminosity densities of the Universe, and show that it is small at all redshifts, rising from $\\sim1-5\\%$ at $z\\sim0$ to a roughly constant $\\sim4-10\\%$ at $z>1$. This agrees well with recent attempts to estimate the contribution to the SFR density at $z=0-1$ specifically from merger-induced starbursts, using either pair or morphologically-selected samples \\citep{jogee:merger.density.08, robaina:2009.sf.fraction.from.mergers}.\\footnote{Note that it is important here to distinguish estimates of the SFR {\\em induced} by mergers, i.e.\\ that above what some control population would exhibit, from that simply in ongoing/identifiable mergers (since many criteria identify mergers for a timescale $\\sim$Gyr, much longer than the burst timescale). For example, a merger fraction of $10\\%$ would imply at least $10\\%$ of star formation in ongoing mergers, even if there were no starbursts and those systems were only forming stars at the same rate as they would in isolation.} Given the completely independent nature of the constraints, and significant uncertainties involved in both, the agreement is good. The small value is what is expected, given our previous determination that the typical burst mass is just $\\sim10\\%$ in $\\sim L_{\\ast}$ spheroids (the galaxies that dominate the stellar mass density). But it clearly rules out merger-induced bursts driving the SFR density evolution of the Universe. At the highest redshifts $z\\gtrsim2$, we can put strict upper limits on starburst intensities, based on the maximum stellar mass remaining at high densities at $z=0$, and find some tension between these and estimated number counts of sub-millimeter galaxies from \\citet{chapman:submm.lfs}. This implies that some change may be necessary in either the number counts themselves, the bolometric corrections used to convert these observations to total IR luminosities, or the stellar IMF used to convert between SFR and IR luminosity. However, the observations remain considerably uncertain, with the bolometric corrections relying sensitively on assumed dust temperatures, and the number counts subject to significant cosmic variance \\citep[see e.g.][]{austermann:2009.aztec.submm.source.counts}. More observations, at new wavelengths and in larger, independent fields, are needed to resolve these discrepancies. We compare our constraints on these histories with recent predictions from galaxy formation models and simulations, and find reasonably good agreement. Both exhibit similar tension with estimates of the high-redshift, extremely luminous number densities. However, the models are able to match the inferred number densities presented here {\\em without} any change to the stellar IMF or requiring other exotic physics. The systematic uncertainties in the models, especially at high redshift, are large, however, so the empirical constraints presented here provide a powerful new means of constraining the models and their input parameters. For example, if models are constrained via e.g.\\ a halo occupation approach or otherwise, so as to match the observed merger fractions of galaxies as a function of redshift, then these numbers become relatively large ($>10\\%$) at redshifts $z\\gtrsim2$ \\citep[see e.g.][]{bundy:merger.fraction.new,conselice:mgr.pairs.updated, kartaltepe:pair.fractions,lin:mergers.by.type,bluck:highz.merger.fraction, hopkins:merger.rates,jogee:merger.density.08, bridge:merger.fraction.new.prep}. The most common assumption in many analytic and semi-analytic models is that, in such a major merger, the entire galaxy gas supply is channeled into the starburst. This, however, coupled with the high observed merger fractions (and implied merger rates), would lead to an over-prediction of the burst contribution to the SFR density, relative to our constraints here. For example, the predictions of such a model are given in \\citet{hopkins:merger.lfs}, and rise to $\\sim20-50\\%$ of the SFR density at $z>1-2$. Recently, however, \\citet{hopkins:disk.survival} have pointed out that the processes that lead to angular momentum loss by the gas -- driving bursts in the first place -- rely on the stellar component of merging systems, and so become less efficient as gas fractions increase. Including these physics in the models (as the current generation to which we compare does) leads to an asymptotic maximum contribution to the SFR density similar to that found here, while still giving a good match to observed merger fractions. Despite the simple nature of the assumptions involved in our analysis, we show in high-resolution hydrodynamic simulations that they work well in allowing us to recover the star formation histories of bursts (even where the simulated system is much more complex). Especially in a statistical sense, our numerical experiments suggest that this approach is robust to a variety of detailed deviations from our idealized assumptions. Testing this (even in a few objects) via high-resolution reconstruction of stellar populations in bursts, directly from their observed spatially-resolved SEDs, would provide a powerful test of this. Insofar as we have compared with different simulations and models, the dominant uncertainty in our analysis is the nature -- in particular the slope, relevant for the extrapolation to high gas surface densities -- of the \\citet{kennicutt98} law. We have considered a range of slopes suggested by different observations. Although they yield similar behavior at moderate and low SFRs, after approximately $\\sim10^{8}$\\,yr from the beginning of a burst, the predicted behavior at very early times in the bursts is quite different. A larger \\citet{kennicutt98} relation index implies higher peak SFRs and more sharply peaked starbursts, increasing the predicted number counts of the most luminous sources and making it relatively more easy for models to account for the most extreme star-forming systems. It therefore remains of considerable importance for observations to probe both gas density measurements and full SFR indicators in extreme systems, at low and high redshifts. Finally, we note that our analysis is only possible because, as indicated by basic dynamics, simulations, and observations of stellar populations, starburst components of spheroids were formed in dissipational (i.e.\\ gas-rich, at their centers), rapid star forming events. The dissipationless ``envelopes'' surrounding the central, dense components in spheroids were not formed in such a manner (again indicated by both their structural and kinematic properties, simulations of their formation, and stellar population observations). Rather, they represent the debris of stars from disks which were formed pre-merger, and assembled (and violently relaxed) dissipationlessly. These stars were formed over extended periods of time, with new gas accretion onto the disk fueling new or continuous star formation \\citep[e.g.][]{keres:cooling.clumps.from.broken.filaments}, as opposed to a single massive inflow. Especially in the most massive systems, they are also assembled from multiple systems, via e.g.\\ minor mergers contributing tidal material to the extended ``wings'' well-known in massive galaxies. As such, there is not a straightforward means to invert their surface stellar mass density profiles to obtain their star formation history. Such constraints will depend on other methods, such as e.g.\\ direct stellar population analysis. It should be borne in mind that this represents $\\sim90\\%$ of the mass in most spheroids, and so understanding star formation in disks remains critical to understanding the origin of stellar populations in spheroids." }, "0910/0910.1472_arXiv.txt": { "abstract": "{}{We compile and analyze an extended database of light curve parameters scattered in the literature to search for correlations and study physical properties, including internal structure constraints.} {We analyze a vast light curve database by obtaining mean rotational properties of the entire sample, determining the spin frequency distribution and comparing those data with a simple model based on hydrostatic equilibrium.} {For the rotation periods, the mean value obtained is 6.95 h for the whole sample, 6.88 h for the Trans-neptunian objects (TNOs) alone and 6.75 h for the Centaurs. From Maxwellian fits to the rotational frequencies distribution the mean rotation rates are 7.35 h for the entire sample, 7.71 h for the TNOs alone and 8.95 h for the Centaurs. These results are obtained by taking into account the criteria of considering a single-peak light curve for objects with amplitudes lower than 0.15 mag and a double-peak light curve for objects with variability $>$0.15mag. We investigate the effect of using different values other than 0.15mag for the transition threshold from albedo-caused light curves to shape-caused light curves. The best Maxwellian fits were obtained with the threshold between 0.10 and 0.15 mag. The mean light-curve amplitude for the entire sample is 0.26 mag, 0.25 mag for TNOs only, and 0.26 mag for the Centaurs. The Period versus B-V color shows a correlation that suggests that objects with shorter rotation periods may have suffered more collisions than objects with larger ones. The amplitude versus H$_v{}$ correlation clearly indicates that the smaller (and collisionally evolved) objects are more elongated than the bigger ones.} {From the model results, it appears that hydrostatic equilibrium can explain the statistical results of almost the entire sample, which means hydrostatic equilibrium is probably reached by almost all TNOs in the H range [-1,7]. This implies that for plausible albedos of 0.04 to 0.20, objects with diameters from 300 km to even 100 km would likely be in equilibrium. Thus, the great majority of objects would qualify as being dwarf planets because they would meet the hydrostatic equilibrium condition. The best model density corresponds to 1100 kg/m$^3$.} ", "introduction": "The belt of remnant planetesimals in the so-called Kuiper belt, whose existence was confirmed observationally in 1992 \\citep{Jewitt1993}, is an important source of knowledge about the formation and early evolution of the Solar System. Features in the orbital distribution of the belt can provide important information about the early dynamical processes such as planetary migration, which was merely untested theories around 25 years ago \\citep{Fernandez1984} that now can be regarded as being well proven. The detailed architecture of the Kuiper belt might also be compatible with a very unstable phase when Jupiter and Saturn entered into resonance \\citep{Tsiganis2005, Morbidelli2005, Gomes2005} around 800Myr after the Solar System formation. Although this particular dynamical scenario is still disputed, there is no question that the architecture of the transneptunian belt can offer invaluable clues about the dynamical mechanisms acting in the early Solar System. In addition to dynamical inferences from the orbital characteristics of the TNOs, much can be learnt about the physical processes that took place in the solar nebula by means of the study of the physical properties of TNOs and related groups of objects. Since the TNOs are understood to be mixtures of ices and rocks from which the Centaurs and ultimately the Jupiter family comets (JFCs) originate \\citep{Jewitt2004b}, ground and space-based observations of all these populations can give us clues about the physics of the outer Solar System. From the ``Nice\" dynamical model \\citep{Tsiganis2005, Morbidelli2005, Gomes2005}, it is also likely that the Jupiter trojans are closely related to the TNOs \\citep{Morbidelli2005}, but this is not yet well established. These objects are regarded as the least evolved objects in the Solar System, but they have experienced many different types of physical processes that have altered them considerably. These processes range from collisions, to space weathering. The collisional processes have probably left clear imprints in both the spin distribution and in the light-curve amplitude distribution of TNOs and Centaurs. The currently accepted theory based on the study of \\cite{DavisFarinella1997} on collisional evolution of the TNOs is that objects of sizes larger than 100 km are very slightly collisionally evolved, whereas the smaller fragments have experienced more collisional events and hence should have their rotations and shapes more significantly altered. This picture is also consistent with newer collisional evolution models \\citep{Benavidez2009}. However, detailed modelling of the evolution of the spin rates related to collisional models has not yet been carried out for the transneptunian belt. Due to the dedicated efforts of two main groups studying short-term variability \\citep{Gutierrez2001, Ortiz2003a, Ortiz2003b, Ortiz2004, Ortiz2006, Ortiz2007, Moullet2008, Duffard2008,Sheppard2007, Sheppard-Jewitt2002, Sheppard2000} and occasional contributions by other authors, a database of about 40 rotation periods and 80 light-curve amplitudes could be extracted from the literature in 2008. Thus, we now have sufficient numbers of TNOs light curves to start analyzing them from a statistical point of view. In a review by \\cite{Sheppard2008}, those data were compiled and some conclusions were outlined. After that compilation, a new set of light curves from 29 TNOs and Centaurs became available \\citep{Thirouin2009} and those authors also presented evidence of possible biases that can be present in the light curve database gathered from the literature. We also developed a simple model to interpret rotation rates and light-curve amplitudes together, to obtain information about densities and tensile strengths. Instead of analyzing each body separately and trying to retrieve information about its density and cohesion by using its spin rate and light-curve amplitude (and an assumed viewing angle), we address the mean internal properties by analyzing the population as a whole. In other words, we developed a model that can simultaneously interpret the distribution of spin rates and the distribution of light-curve amplitudes. In this paper, light curve parameters of a vast sample are compiled and analyzed. We focus on the analysis of the period and shape distribution presented in Sect. 2. Then, with the current data, we complete a correlation analysis with several orbital parameters in Sect. 3. Internal properties of the TNOs are derived by applying a simple model in Sect. 4, a discussion of the results is presented in Sect. 5. Finally, the conclusions are presented in Sect. 6. ", "conclusions": "We have analyzed the light-curve data of \\cite{Thirouin2009} and others in the literature to derive light-curve parameters, which have been compiled here. We have presented the most complete distribution of rotation periods and amplitudes so far. For the rotation periods, the mean value obtained by a Maxwellian fit in rotational frequencies is 7.71 hours (3.11 cycles/day) considering only the TNOs, 8.95 hours (2.68 cycles/day) considering only Centaurs, and 7.35 hours (3.19 cycles/day) considering all the sample altogether. These results were obtained by taking into account the criteria of considering a single peak light curve for those objects with amplitudes lower than 0.15 mag and a double peak light curve for the more elongated ones. This mean rotational period is slightly longer than the mean value for the largest Main Belt asteroids, 6.9 hours \\citep{Binzel1989}, but we state that this value may be affected by observational biases. It is important to stress that in photometry devoted to the study of the rotational properties of minor bodies, there exists and important observational bias towards shorter periods and higher amplitude light curves. How this bias can be estimated remains an open problem, but it certainly depends on the number of objects for which a period is obtained. Some steps in this direction were already taken by \\cite{Thirouin2009}. Concerning other biases, if observational time for larger telescopes is granted then smaller TNOs will be accessible for the determination of the light curves, and the sample will not only increase in size but the bias in size will become less important. One of the most important correlations found in this work is the light curve amplitude and absolute magnitude of all the sample of KBOs. This correlation is not a surprise and is probably related to the most collisionally evolved small population, which should be more elongated objects. On the other hand, the less collisionally evolved population (larger and then brighter objects) is the more numerous one in our sample ($\\sim$60\\%), so we caution the reader that some of our results could be somewhat biased. Another interesting correlation is the rotational period and B-V color, which could indicate the age of the surface (younger surfaces being bluer). This correlation appears to be consistent with the idea that the more collisionally evolved objects spin faster and have bluer surfaces. A simple shape model has been developed that is able to reproduce some results on the relative percentage of presence of MacLaurin/Jacobian figures, when assuming hydrostatic equilibrium. In the entire sample, less than 12\\% of the objects in hydrostatic equilibrium have a Jacobi shape. Most of the objects have MacLaurin figures that have small light curve amplitudes (because of albedo marks). The model that best fits the observations requires densities between 1000 to 1200 kg/m$^3$. We must mention that as was highlighted in \\cite{Thirouin2009} there is an observational bias in the percentage of low amplitude light curves, which means that perhaps 75\\% of the objects have low light curve amplitudes. This implies that the mean density is likely somewhat higher, around 1500 kg/m$^3$ (see Fig. \\ref{fig08}). For plausible albedos of 0.04 to 0.20, the absolute magnitude range H=[-1,7] includes objects with diameters as small as $\\sim$120 km. Those objects would also be in hydrostatic equilibrium. We end with the provocative thought that the great majority of observed TNOs would qualify to be dwarf planets, because they appear to meet the hydrostatic equilibrium condition." }, "0910/0910.0018_arXiv.txt": { "abstract": "The low-mass X-ray binary microquasar GRO~J1655--40 is observed to have a misalignment between the jets and the binary orbital plane. Since the current black hole spin axis is likely to be parallel to the jets, this implies a misalignment between the spin axis of the black hole and the binary orbital plane. It is likely the black holes formed with an asymetric supernova which caused the orbital axis to misalign with the spin of the stars. We ask whether the null hypothesis that the supernova explosion did not affect the spin axis of the black hole can be ruled out by what can be deduced about the properties of the explosion from the known system parameters. We find that this null hypothesis cannot be disproved but we find that the most likely requirements to form the system include a small natal black hole kick (of a few tens of $\\rm km\\, s^{-1}$) and a relatively wide pre-supernova binary. In such cases the observed close binary system could have formed by tidal circularisation without a common envelope phase. ", "introduction": "It has long been known that neutron stars have much greater space velocities than their likely progenitors \\citep{GO70} and it is now widely accepted that that this is because they acquire large velocity kicks in the supernova explosions in which they form \\citep{S70,S78}. The reasons for these kicks is still a matter for debate with the leading candidates being asymmetric neutrino emission and/or asymmetric mass release during the supernova explosion and core collapse \\citep{BP95,Pod02}. We address here the question of the extent to which similar kicks may be present when black holes form. Because the material forming the black holes passes through the event horizon it is quite possible that few neutrinos can escape \\citep{gour93} so that the black hole could form with little or no natal kick. On the other hand, asymmetric collapse to form a black hole might lead to copious emission of gravitational waves \\citep{bonnell95, kobay03}. In addition, \\cite{LL94} note that in a binary system a black hole can form by accretion of matter on to a neutron star. In that case the kick given to the original neutron star would appear as a kick given to the current black hole. Evidence that black holes are indeed kicked comes from the work of \\cite{jonker04}. They looked at the out-of-plane distributions of low mass X-ray binaries and neutron stars. They found no significant difference between the two, leading to the conclusion that black holes are subject to similar kicks at their formation. In this paper we accept the evidence that stellar black holes do indeed acquire a velocity kick when they form, but we inquire further into the nature of the kick. In particular we ask whether the mechanism which gives rise to the kick might also give rise to a misalignment of the black hole spin axis with the original spin axis of the star from which it formed. The black hole progenitor star most likely had its spin aligned with the binary orbit before the supernova. A simple spherical collapse would preserve the spin axis, but a more complicated collapse might not. We consider here the simple null hypothesis that the spin axis remains unaltered by the kick process and test the extent to which this might be contradicted by the evidence. We focus on the microquasar GRO J1655-40. As we discuss in Section~\\ref{GRO}, there is considerable information for this system, about the size of the natal velocity kick and also on the misalignment between the current black hole spin and the binary orbital axis. In Section~\\ref{model} we outline the dynamics of natal kicks and their implications for spin/orbital misalignment and in Section~\\ref{apply} apply these results to GRO J1655-40. We discuss our results in Section~\\ref{conclude}. ", "conclusions": "\\label{conclude} We have considered what can be learned about the natal kick acquired by a black hole in a supernova by considering the system GRO~1655--40. In line with the analysis of \\cite{Willems05} we make use of kinematic data, but we add the additional constraint that the current black hole spin is not aligned with the orbital rotation. Of course, if the supernova explosion and collapse process in which the black hole is formed gives not just a linear impulse, but also angular momentum, to the hole, then the current misalignment is just a measure of the added angular momentum. The task we set ourselves is to ask if it is possible, given the various constraints, to rule out the possibility that the natal kick (if any) added just linear, and not angular, momentum. To simplify the analysis we have fixed the masses of the stars before and after the supernova. To be fully consistent we could allow these masses to vary too and assign probabilities based on the likelihood of a given set of parameters. However one set of masses proved to be sufficient to not rule out the null hypothesis. Our main results are summarised in Table~\\ref{xray}. We have considered randomly oriented natal kicks with Maxwellian velocity distributions of $265$ and $26.5\\,\\rm km \\,s^{-1}$ for various separations of the pre-supernova binary, parametrised in terms of its relative orbital velocity $v_{\\rm orb}$. Column~2 gives the probability $P_1$ that a bound system with the appropriate properties can be formed, ignoring any information about misalignment. It is immediately evident, that for low-velocity kicks the probability of forming the system is high, being typically around a few tens of percent \\citep[compare][]{Willems05}. The probabilities are lower for higher-velocity kicks, being typically around a percent. That is only around one in a hundred such pre-supernova systems would end up looking like GRO~1655--40. This is still probably not unreasonable. The remaining columns in Table~\\ref{xray} show the probabilities, given $P_1$ that the additional constraint of the misalignment angle $i$ can be satisfied, given our null hypothesis. It is evident that we are not able to rule out this null hypothesis, and so it is quite possible that the black hole formation process does not affect the angular as well as the linear momentum of the resulting black hole. It is worth noting, however, that even without the alignment information, it is easier to form the observed system with a low-velocity kick. Moreover, if the formation process did just impart linear momentum, then, with account for the fact that the likely alignment timescale is comparable to the age of the system since the formation of the hole \\citep{martin2008} so that the initial misalignment angle $i$ would have been in the range $20^\\circ - 40^\\circ$, then it is evident that there is a strong preference for the pre-supernova system to have be fairly wide. Note that $v_{\\rm orb}=50\\,\\rm km\\,s^{-1}$ corresponds to a binary separation of around $4.6\\,\\rm AU$ before the supernova. It is interesting that such a wide system could have avoided the traditional common envelope required to shrink the orbit \\citep{verbunt96} We also note, that if the current misalignment angle were large (say, $i > 40^\\circ$) then it would become increasingly difficult to satisfy the constraints for GRO~1655--40. In this respect we note that V4641 Sgr is a microquasar similar in many respects to GRO~1655--40 and it has a current misalignment angle of $i \\approx 55^\\circ$ \\citep{O01}. If further observations were able to provide a constraint on its post-supernova systemic velocity, $v_{\\rm sys}$ then it might be able to comment more securely on the null hypotheses. For example if it could be shown to have a large space $v_{\\rm sys}$ then satisfying the null hypothesis might be problematic." }, "0910/0910.5910_arXiv.txt": { "abstract": "{Neutral atomic hydrogen (\\hi) traces the interstellar medium (ISM) over a broad range of physical conditions. Its 21-cm emission line is a key probe of the structure and dynamics of galaxies. This line comprises very different temperature and density regimes on all scales, from tens of astronomical units to kiloparsecs. \\hi is the key element to study the evolution of the ISM in detail. To understand the physics of the ISM and to analyze the interplay between different phases it is mandatory to cover observationally a broad range of scales. But large scale imaging of galaxies, resolving at the same time all these scales is difficult; spatial resolution, as well as sensitivity and the field of view are currently rather limited. The observational situation is much more favorable if we consider our own galaxy. Two major all sky 21-cm line surveys of the Milky Way will become available soon. The Galactic All Sky Survey (GASS) obtained with the Parkes 64-m telescope\\thanks{The Parkes Radio Telescope is part of the Australia Telescope which is funded by the Commonwealth of Australia for operation as a National Facility managed by CSIRO}~ for the southern hemisphere with a resolution of 16 arcmin is close to completion. The northern extension, the Effelsberg Bonn HI Survey (EBHIS)\\thanks{Based on observations with the 100-m telescope of the MPIfR (Max-Planck-Institut f\\\"ur Radioastronomie) at Effelsberg}~ with 9 arcmin resolution, will be available in 2010/2011; we refer to the talks by Kerp and Winkel. Here we discuss briefly the GASS and demonstrate the unprecedented quality of this survey. The Galactic single dish 21-cm line surveys prepare the ground for future high resolution imaging of the Galactic \\hi distribution. Using the available short spacing informaton, the Australian Square Kilometre Array Pathfinder (ASKAP) will be capable to generate a truly panoramic view of the Milky Way HI gas distribution with arcsecond resolution for all declinations $ < 30\\deg$. Data from the Widefield ASKAP L-band Legacy All-Sky Blind Survey (WALLABY, see talk by Staveley-Smith) can be used to generate high resolution all sky maps. In comparison to the currently available interferometric International Galactic Plane Surveys (IGPS) the sensitivity will improve by a factor of 10. Most important is the all sky coverage which will overcome the rather limited spatial coverage of a few degrees around the Galactic plane for the IGPS.} \\FullConference{Panoramic Radio Astronomy: Wide-field 1-2 GHz research on galaxy evolution\\\\ June 2-5 2009\\\\ Groningen, the Netherlands} \\begin{document} ", "introduction": " ", "conclusions": "" }, "0910/0910.5121_arXiv.txt": { "abstract": " ", "introduction": "Large redshift surveys are typically completed by observing with a Multi Object Spectrograph (MOS), obtaining spectra for many hundreds of sources simultaneously over large fields of view. The problem of how to optimise observing strategies to target sources distributed over some survey area with a given MOS, defining a field of view and number of simultaneous targets, falls into the ``area packing\" class of problems. Much work outside of astronomy has been devoted to such problems \\citep{megi84} which are usually intractable in a formal, provably-optimal, sense. In the case of the Anglo-Australian Telescope's (AAT) largest survey to date, the 2 degree Field Galaxy Redshift Survey \\citep[2dFGRS,][]{coll01}, the survey was created in a manner that minimised field overlaps in order to maximise area (the target magnitude limit being $b_{j}=19.45$). This obviously had an impact on the target completeness, and the observations had to be weighted in order to account for the local levels of incompleteness. At the other extreme is the 6 degree Field Galaxy Survey \\citep[6dFGS,][]{jone04} that aimed for high levels of completeness within the local universe. In this case the filamentary structures (i.e.\\ non-uniform overdensities) present on small scales necessitate extremely non-uniform tile coverage and potentially large amounts of overlap among tiles, target densities varying from 6 to 30 galaxies per deg$^{2}$. Hence the optimal strategy for tiling is closely linked to the scientific objectives of the survey, and a generic approach will not be appropriate for all requirements. Fibre fed MOS instruments typically have a circular field of view (FOV), as seen for example in the 2 degree Field \\citep[2dF,][]{lewi02}, 6 degree Field \\citep[6dF,][]{jone04}, Sloan Digital Sky Survey (SDSS) Spectrograph \\citep{york00}, Hectospec \\citep{fabr05} and Hydra \\citep{bard93}. Also typical is for survey regions to be rectangular in spherical coordinate geometry: recent examples include the 2dFGRS, Sloan Digital Sky Survey \\citep[SDSS,][]{abaz09} and Millennium Galaxy Catalogue \\citep[MGC,][]{lisk03,driv05}. This latter commonality is due to a number of allying factors: imaging CCDs used for input catalogues are almost always rectangular\\footnote{The use of GALEX in WiggleZ \\citep{glaz07} is a rare counter-example.} and survey boundaries and volumes are easier to consider when using spherical coordinate derived edges. Packing a shape best described in spherical coordinates into a Cartesian defined region is a non trivial task, and many approaches have been used in redshift surveys. Such packing problems are of wider mathematical interest because no provably optimal and rapid technique has yet been discovered \\citep{megi84}. Instead every large survey tailors a tiling method in line with specific survey goals using a heuristic method. In this sense a heuristic method is one informed by knowledge of the problem at hand, the hope being the solution is not much worse than optimal. On top of the generic problem of efficient tile packing, spectroscopic surveys also have to contend with extremely non-uniform and complex selection functions within the tiles themselves. The major cause for the non-uniformity is object exclusion, either due to fibre collisions or slit overlaps. In the case of 2dFGRS an approach close to hexagonal packing was used, where slight perturbations were made to a purely hexagonal grid of tile centres in order to better sample object densities. Since this survey was almost single pass (there was $\\sim 30\\%$ tile overlap), low completeness fields were not uncommon, an effect that was statistically adjusted for with observational weights. However, in the densest fields some targets will not have redshifts, and galaxy group assignments will not be as secure as in highly complete regions. The downside of such a regular approach is that all multi-fibre spectrographs will have structure or bias in their assignments and thus power will be added to (or removed from) certain frequencies in tangential modes of the power spectrum. The distribution of fibres is not only driven by the algorithm used to place them, but also the physical limitations of the instrument. Typically a fibre fed MOS is designed with fibres around the circumference in such a way that all fibres can reach the centre and few can reach locations at the edge, a scenario that makes radially-dependent targeting distributions inevitable. Even with the newest simulated annealing (SANN) algorithms available for AAOmega, radial assignment dependencies within each 2dF pointing exist \\citep{misz06}. It is obviously important to try to compensate for such biases in any work that is concerned with clustering and structure, such as Galaxy And Mass Assembly \\citep[GAMA,][]{driv09}, the latest large survey to use AAOmega on the AAT. The spectroscopic element of SDSS \\citep{blan03} used a heuristic algorithm that attempted to find an acceptable solution of a perturbed uniform grid of tiles, much like 2dFGRS. The algorithm aimed to utilise 90\\% of the 600 available fibres on each tile, and similar to 2dFGRS the SDSS's median tile coverage for an object was 1 (both achieved a target density $\\sim$100 galaxies per square degree). Minimum fibre spacings are $55''$ for the SDSS spectrograph, larger than the $40''$ distance for 2dF, thus an obvious limitation of SDSS is the full targeting and unbiased analysis of close pairs (a key science objective for GAMA, discussed in detail below). Of recent surveys, the VIMOS VLT Deep Survey (VVDS from here) utilised the simplest approach to tiling \\citep{bott05}. Effectively it placed tile centres on a fixed square grid with diagonal offsets used for the deeper component of the survey. Such an approach is possible when using VIMOS because of its mask-based grism spectrograph, giving it a square FOV better suited to tiling a square CCD photometric survey. The VVDS does not suffer from any radial selection bias, but due to the constraints imposed by slits cut into each mask it does possess complex selection effects such as the tendency to target a uniform spread of targets; highly clustered regions are hard to target since the slits necessarily avoid each other. Further complicating matters is a partially radial completeness bias, evident in the spectroscopic masks created for zCOSMOS \\citep{knob09}. Whilst an interesting survey to note, such a survey design is not trivial to create with any of the fibre based multi object spectrographs discussed due to their circular FOV, and the complex radial bias this introduces. Simulated annealing solutions of the tiling problem have been utilised in large area surveys with large amounts of structure present, most notably by the 6dFGS \\citep{camp04}. Simulated annealing is a popular approach for many algorithmically insolvable problems and is, strictly speaking, a metaheuristic solution (i.e.\\ choices have to be made about the element to be optimised and also the method of optimising). In simple terms the user must pick something to be minimised (or maximised), such as the total number of objects not assigned to a fibre after tiling the whole survey region. The user must also give the SANN algorithm variables to perturb (most obviously the right ascension and declination of the tile centres), and a rate at which it `cools' towards a solution. Typically these perturbations become smaller as the solution improves, and eventually an acceptable set of tile positions should be found. Packing problems lend themselves well to SANN since they can be tuned to find acceptable solutions rapidly, but they are non-deterministic algorithms (unlike the other heuristic approaches discussed) and are neither provably optimal nor stable (i.e.\\ small variations to the problem to be optimised can produce radically different results). In the case of the 6dFGS, SANN is obviously much more effective than any sort of regular tiling because the projected target densities vary significantly and the survey area is large. The use of SANN reduces the number of sparsely populated fields and better samples overdensities where fields would be full. Added to the complexities of these different approaches are the observational limitations for any survey as well as its scientific priorities. It will not be the case that all fields are equally observable in a large area survey (e.g.\\ varying rising and setting time as a function of RA), but in a sufficiently small area survey it will often be the case that all parts of the survey field are effectively as observable as each other. Also, the end point of the survey will often be an unknown (i.e.\\ weather dependent), so in many applications it is advantageous for the survey to be in a useable state as quickly as possible. With these extra considerations in mind, the philosophy that was applied to tiling GAMA was one where each tile would in some sense be the next most optimal tile, and every subsequent tile should make a significant impact towards achieving the GAMA survey requirements. The GAMA redshift survey is one component of the multi-band GAMA survey project, and is the latest large survey to use the AAT's MOS facility. In this paper we explore the problems of tiling specifically for the GAMA survey, with the possibility of using the approaches discussed in future redshift surveys with characteristics in common with GAMA. In section 2 we outline the GAMA survey, and how the scientific goals for the project translate into survey requirements that our tiling algorithm must achieve. In section 3 we discuss in detail the different options to tiling that are appropriate for GAMA. In section 4 we apply the two most likely candidates for the tiling algorithm to the GAMA survey as it was left at the end of year 1, allowing quantitative judgements of the different approaches to be made. In section 5 we apply our chosen tiling algorithm to the data and present the state of the survey after year 2. Finally, conservative predictions are made for the state of the survey after year 3 observations based on tiling simulations. ", "conclusions": "This work has demonstrated that a greedy approach to tiling proves to be extremely successful in densely packed surveys such as GAMA. By aggressively targeting under-densities with each field used, high levels of spatial completeness should be a reality for each GAMA region by the end of the third year of observations. In the meantime we allow for a simple mechanism to feed redshift failures back in, and by prioritising highly clustered regions we obtain both a large number of close pairs, and guarantee we are not left with difficult pockets of galaxies in the final stages of the survey. Further still, by utilising every non-main survey fibre on deeper targets, we ensure efficient use of the 2dF instrument, and make a head-start on any future extended redshift surveys in the GAMA regions." }, "0910/0910.0704_arXiv.txt": { "abstract": "Infrared spectroscopy of the \\ha{} emission lines of a sub-sample of 19 high-redshift (0.8 $< z <$ 2.3) Molonglo quasars, selected at 408 MHz, is presented. These emission lines are fitted with composite models of broad and narrow emission, which include combinations of classical broad-line regions of fast-moving gas clouds lying outside the quasar nucleus, and/or a theoretical model of emission from an optically-thick, flattened, rotating accretion disk, with velocity shifts allowed between the components. All bar one of the nineteen sources are found to have emission consistent with the presence of an optically-emitting accretion disk, with the exception appearing to display complex emission including at least three broad components. Ten of the quasars have strong Bayesian evidence for broad-line emission arising from an accretion disk together with a standard broad-line region, selected in preference to a model with two simple broad lines. Thus the best explanation for the complexity required to fit the broad \\ha{} lines in this sample is optical emission from an accretion disk in addition a region of fast-moving clouds. We derive estimates of the angle between the rotation axis of the accretion disk and the line of sight. Deprojecting radio sources on the assumption of jets emerging perpendicular to the accretion disk gives rough agreement with expectations of radio source models. The distribution in disk angles is broadly consistent with models in which a Doppler boosted core contributes to the chances of observing a source at low inclination to the line of sight, and in which the radio jets expand at constant speed up to a size of $\\sim 1$ Mpc. A weak correlation is found between the accretion disk angle and the logarithm of the low-frequency radio luminosity. This is direct, albeit tenuous, evidence for the receding torus model first suggested by \\citet{lawrence91} in which the opening angle of the torus widens with increasing radio luminosity. The highest accretion disk angle measured is 48$\\degree$, consistent with the opening angle predicted for radio-luminous sources. Velocity shifts of the broad \\ha{} components are analysed and the results found to be consistent with a two-component model comprising one single-peaked broad line emitted at the same redshift as the narrow lines, and emission from an accretion disk which appears to be preferentially redshifted with respect to the narrow lines for high-redshift sources and blueshifted relative to the narrow lines for low-redshift sources. An additional analysis is performed in which the disk emission is fixed at the redshift of the narrow-line region; although only two quasars show a robust change in fitted angle, the radio luminosity -- disk angle correlation falls sharply in probability, and so is strongly model dependent in this sample. ", "introduction": "\\subsection{Orientation effects} \\label{sec:orientation} Optical emission spectra of Active Galactic Nuclei (AGN) matched in radio and optical luminosity are now firmly believed to be strongly influenced by orientation effects, with only small underlying differences in the sources themselves. There are two separate orientation-dependent effects which alter the optical spectra of AGN. The Type 1/Type 2 classification of AGN is made according to the presence of broad emission lines. Type 2 AGN possess only narrow emission lines, $\\lesssim 2000$ \\kms{} \\citep[e.g.][]{peterson97} and weak non-stellar continuum emission. Type 1 AGN have broad emission lines of $\\sim 2000$ -- $20000$ \\kms, and strong non-stellar continuum emission, in addition to narrow lines similar to the Type 2 sources. An explanation for this disparity grew from the discoveries of broad emission lines seen in polarised light from Type 2 Seyfert sources \\citep[e.g.][]{antonucci85}, which suggested orientation-dependent obscuration caused by an intervening screen of matter, such as a dusty molecular torus \\citep{krolik86}. It has recently become clear that obscuration by dust in starbursting galaxies can also be responsible for concealing Type 2 AGN \\citep[e.g.][]{martinez05}. The second orientation-dependent effect arises from the relativistic motion of the plasma in the radio jets. \\citet{scheuer79} first suggested that viewing a radio source with opposing relativistic jets would cause a large contrast in the luminosities of the approaching and receding jets, and that objects with jet axes close to the line of sight would be seen more often due to Doppler boosting of the core. \\citet{orr82} made the connection between Doppler boosting of the core, and a measure of quasar orientation from the core-to-lobe radio flux density ratio; this allowed them to unify the ``core-dominated'' quasars with flat optical spectra, viewed at angles close to the line of sight, with the ``steep-spectrum'' or radio-lobe-dominated quasars viewed at larger angles. \\citet{wills86} linked the radio properties to the optical properties by discovering an anticorrelation between the core-to-lobe radio flux ratio and the width of the \\hb{} line, interpreting this connection as the result of beaming of radio emission from a jet emerging along the rotation axis of an accretion disk; the broad \\hb{} lines arise from the accretion disk with a width correlated with the angle between the line of sight and the disk axis. The two optical schemes were reconciled by \\citet{barthel89}, who gave a consistent picture in which FRII narrow line radio galaxies, steep-spectrum quasars and flat-spectrum quasars are all drawn from the same parent population, but viewed at decreasing angles to the line of sight. A review of these so-called unified schemes for AGN can be found in \\citet{urry95} or \\citet{antonucci93}. A solid understanding of how quasar emission lines arise, and how they are affected by the AGN environment along different sight lines, is not only an interesting study in terms of quasar structure, but is also vital in order to disentangle orientation effects from large-scale AGN surveys which enable the study of cosmic evolution. \\subsection{Accretion disks} \\label{sec:disks} There is a growing body of evidence that AGN are powered by accretion of gas and dust onto supermassive black holes, and this is now the accepted paradigm. As the black hole feeds on the surrounding material, it is expected that this will form an accretion disk of infalling matter \\citep{shakura73}. The current theory is that accretion disks have two parts: a puffed-up, X-ray-emitting inner disk, and a flattened, outer disk which gives rise to broad optical emission lines. \\citet{collin80} first suggested that a thickened inner accretion disk, shielding and reprocessing the hard X-ray emission from the black hole, gives rise to the optical low-ionisation \\feii{} emission seen in Type 1 Seyferts, from the atmosphere above the geometrically-thin outer part of the disk. \\citet{rees82} postulated that a geometrically- and optically-thick ion-supported torus surrounds the supermassive black hole at the centre of a radio galaxy, collimating the emerging radio jets. \\citet{tanaka95} observed a broad, asymmetric iron K$\\alpha$ line consistent with emission from a disk of this description, situated between $\\sim 3$ -- $10 \\rg$ (where $\\rg = G M / \\mathrm{c}^2$ is the gravitational radius) from the nucleus of the AGN. \\citet{filippenko88} reviewed the different lines of evidence from optical and UV data that flattened, extended accretion disks fuel the central black holes of galaxies. For example, \\citet{baldwin77} recorded the anticorrelation of the UV continuum luminosity with the equivalent widths of broad \\civ{} emission (the ``Baldwin Effect''). Both this observation and the ``big blue bump'' of excess UV continuum emission found by \\citet{malkan82} may be explained by emission from an optically-thick, geometrically-thin accretion disk \\citep{netzer85}. The strongest direct evidence for this scenario is the presence of double-peaked, low-ionisation optical emission lines seen in some AGN, which arise from the Doppler effect acting on the emission from rotating material in the outer accretion disk (e.g. \\citet{chen89}, \\citet{perez88}). The thin, optically-emitting disk must be illuminated by some mechanism. The photoionising flux may originate either from the central non-thermal source, or from the inner, X-ray-emitting disk \\citep{collin80}; and the radiation may illuminate the outer disk directly, or be scattered from a highly-ionised diffuse medium above the outer disk \\citep{chen89b}. Optical double-peaked lines have to date only been found in a relatively low percentage of radio-loud AGN ($\\sim$ 10$\\%$, see \\citet{eracleous94}). \\citet{strateva03} discovered that radio-quiet quasars also emit double-peaked lines, although these appear to be rarer still: double-peaked emission was seen in $\\sim$ 3$\\%$ of the SDSS AGN, including both radio-loud and radio-quiet sources. Double-peaked profiles are not unique to Balmer lines: \\citet{strateva03} found double-peaked \\mgii{} emission lines in some SDSS AGN. It is not clear why the double-peaked line profiles should be rare, although there are several possibilities: the outer accretion disk may simply be obscured by broad-line-emitting clouds surrounding it; or if the outer accretion disk causes a wind, then the broad lines which arise from this wind are predicted to be single-peaked \\citep{murray97}. In any case, it should not be expected that the low-ionisation broad lines seen in an AGN originate solely from the rotating disk; single-peaked emission may be seen in addition to double-peaked profiles. The accretion disk model used in this paper is taken from \\citet{chen89}, and consists of a thick, hot torus, whose outer edge may reach up to $100\\, \\rg$ from the black hole. Inverse Compton-scattered X-rays from this inner disk illuminate an optically-thick, flattened outer disk \\citep{halpern89}. The thin, circular disk, which produces the double-peaked emission lines, may extend up to $\\sim 10^5\\, \\rg$ from the central engine. The distinctive line profiles are caused by the rotation of the disk, splitting the emission into redshifted receding material and blueshifted approaching material; the blueshifted peak is of higher intensity than the redshifted peak, as a result of Doppler boosting. The chosen model was necessarily simple, to limit the number of free parameters. More complex disks, such as an elliptical disk, which may arise when a single star is disrupted near the black hole \\citep{gurzadyan79}, or a warped disk, thought to occur around rotating black holes \\citep{bachev99}, give rise to a wider range of line profiles, including double-peaked profiles with a redward peak of higher intensity than the blueward peak. \\citet{strateva03} found that assuming all low-ionisation broad-line emission comes from a disk, non-axisymmetric disks would be required in $\\sim 60\\%$ of their SDSS sample, while \\citet{eracleous03} found that in their sample of 106 radio-loud AGN, 20\\% have double-peaked \\ha{} profiles visible to the eye, of which $\\sim 40\\%$ require a model more complex than the circular Keplerian disk. In this paper, a small, but close to complete, sample of radio-loud quasars are analysed to determine if their spectra include emission from circular, planar accretion disks, either as the sole component of broad optical emission, or in combination with a single-peaked broad line. \\subsection{Velocity shifts} \\label{sec:velshifts} It has been generally accepted that the narrow-line region (NLR) of an AGN falls near the systemic redshift. The NLR is extended, and appears to be reasonably independent of viewing angle effects, and so the narrowness of the lines constrains the velocity of this gas to be small. \\citet{heckman81} showed, for a handful of low-$z${} Seyferts and radio galaxies, that the narrow lines had small blueshifts of between $\\sim$ 0 -- 300 \\kms{} relative to the neutral hydrogen emission of the host galaxies. \\citet{vandenberk01} discovered that, for composite spectra created with several thousand Sloan Digital Sky Survey (SDSS) quasar spectra covering a wide redshift range ($0.04 \\lesssim z \\lesssim 4.8$), the narrow lines are blueshifted by small velocities ($\\lesssim 100$ \\kms) which correlate with ionisation potential. The broad-line region (BLR) has a complex structure, with emission line redshifts which vary according to species, implying separate regions of gas (see Figure 6 of \\citet{collin80}). It is thought that the high-ionisation lines (HILs) arise from a compact, spherical region close to the central black hole, while the low-ionisation lines (LILs) are formed in a flattened structure further from the AGN centre, possibly an accretion disk \\citep{krolik91}. \\citet{collin88} suggested that the HILs might be produced by shocks in an outflowing wind. \\citet{gaskell82} demonstrated for a sample of flat-spectrum quasars with z $\\sim$ 0.2 -- 2.3 that the HILs, such as \\ciii, \\civ{} and \\nv{}, are blueshifted by $\\sim 600$ \\kms{} with respect to the LILs, which include \\mgii{}, \\oi{} and the Balmer lines. \\citet{wilkes84} confirmed this shift for high-$z${} quasars ($2 \\lesssim z \\lesssim 3$), finding a slightly higher range of shifts, up to $\\sim$ 1400 \\kms. Blueshifted HIL zones have also been observed by \\citet{espey89}, who found shifts of $\\sim$ 1000 \\kms{} in a small sample of $1.3 \\lesssim z \\lesssim 2.4$ sources, and \\citet{corbin90}, who found shifts exceeding 4000 \\kms{} for luminous, optically-selected sources with $z \\gtrsim 1$. \\citet{richards02} also confirm this trend from studies of $\\sim 800$ quasars with $1.5 \\lesssim z \\lesssim 2.2$ from the SDSS, though they find a wide distribution of velocity shifts of the high-ionisation \\civ{} with respect to the low-ionisation \\mgii, ranging from redshifts of $\\sim$ 500 \\kms{} to blueshifts of over 2000 \\kms, and they take pains to point out that they do not believe it is a line shift so much as a lack of flux in the red wing of the line. \\citet{mcintosh99} found that the HIL zone is at the same redshift as the narrow lines in low-$z${} sources, but for a sample of quasars with $2 \\lesssim z \\lesssim 2.5$, broad \\hb{} had a redshift of $\\sim$ 500 \\kms{} relative to \\oiii. Although the LILs are typically considered to have low velocity shifts, there have been recorded instances of Balmer lines with very high redshifts, e.g. $\\sim$ 2100 \\kms{} for 3C277 \\citep{osterbrock76} and $\\sim$ 2600 \\kms{} for OQ208 \\citep{osterbrock79}. \\section[]{Sample Selection} \\label{sec:selection} A sub-sample of 19 quasars was defined from the Molonglo Quasar Sample \\citep{molonglo3}. The Molonglo sample of radio sources was selected from the Molonglo Reference Catalog (MRC) \\citep{large81}, a 408 MHz survey conducted with the Molonglo Synthesis Telescope which is 99.9\\% complete at 1 Jy. The radio sources were then identified from VLA 1-arcsecond resolution radio images, optical imaging and spectroscopy, the quasars \\citep{molonglo3} being distinguished from the radio galaxies \\citep{molonglo1} by the presence of broad optical emission lines. The Molonglo quasar sample includes all quasars with flux densities \\mbox{$S_{408 \\mathrm{MHz}} > 0.95$ Jy} in a $10\\degree$-wide strip in the Southern sky, \\mbox{$-20\\degree > \\delta > -30\\degree$}, excluding sources with low Galactic latitude \\mbox{($\\mid b \\mid > 20\\degree$)} and also a strip in Galactic R.A. (details in \\citet{molonglo3}) to define a sample of manageable size. It should be noted that as the Molonglo sample was selected at the mid-range frequency of 408 MHz, there are likely to be some sources included in the sample by virtue of their strong radio core emission, and are therefore inclined at small angles to the line of sight; this builds an orientation bias into this survey. The quasar sub-sample was selected on the basis of four criteria: observability during the relevant time period; redshift such that \\ha{} and \\hb{} emission falls in wavelength windows corresponding to regions of high atmospheric transparency ($0.8 < z < 1.0$, $1.5 < z < 1.65$, $2.2 < z < 2.3$); sufficient J-band brightness to be observable in a reasonable integration time; and exclusion of the RA range 14h -- 03h in accordance with scheduling constraints. The J-band magnitude limit is the only factor which adds a significant bias in the sub-sample selection. The objects were selected to be brighter than \\mbox{J $\\sim 18.5$}, and this excluded one source from the sample, MRC0418-288. This source is likely to be reddened or dusty, which means that it has a higher chance of being inclined at a large angle to the line of sight. MRC1256-243 is an extra core-dominated source which was not observed due to scheduling limitations; this source is likely to have been boosted into the sample by virtue of its strong core emission in any case. Throughout the paper, a cosmology of $H_0 = 70$ km s$^{-1}$ Mpc$^{-1}$, $\\Omega_{\\mathrm{M}} = 0.3$ and $\\Omega_{\\Lambda} = 0.7$ is assumed. ", "conclusions": "\\label{sec:conclusions} All bar one of the 19 quasars are best fitted with an emission model which includes more than one component of broad-line emission, in both analyses; there is overwhelming evidence that in this sub-sample of quasars, a simple one-component broad-line emission model is insufficient to describe the physical processes at work. There is strong Bayesian evidence that ten out of nineteen quasars from the analysis in which a shift is permitted between the BLR and NLR possess accretion disks, and all but one are consistent with the emission model of a circular Keplerian accretion disk in addition to a single-peaked, symmetric emission line arising from a separate BLR. The one quasar for which this does not apply appears to have very complex broad emission, consisting of at least three components. If the model does not allow for a shift between the BLR and NLR, strong evidence for disks is seen in eight out of the nineteen sources, and all but three are consistent with the presence of a circular accretion disk. \\citet{eracleous94}, \\citet{eracleous03} and \\citet{strateva03} found that only 3 -- 20$\\%$ of AGN have obvious double-peaked lines (see Section \\ref{sec:disks}), in contrast to the $\\sim$ 84 -- 95\\% of this sample of quasars which were found to be consistent with the presence of a thin, optically-emitting accretion disk, despite only a few of the line profiles appearing obviously double-peaked to the eye. It may be that accretion disk emission has been seen in so few AGN because single-peaked broad emission lines, arising from fast-moving clouds located outside the thin disk or from a wind originating from the disk itself, have a tendency to swamp the disk emission. In these cases, a complex model and high signal-to-noise ratio spectra are required in order to retrieve the accretion disk contribution to the emission. It is therefore essential to extend this analysis to a larger sample of low-radio-frequency-selected quasars; optically-selected quasars are predicted to cover a narrower range in disk angles, as even lightly reddened quasars (at $\\sim 45 \\degree$) are much more likely to fall through the magnitude limit of a survey, and the optical emission may be angle dependent. There is a possible correlation of the fitted disk angles with the projected source size, in agreement with the expected projection effects if the accretion disk is perpendicular to the jet. There are two CSS sources which do not appear to lie on the same relation as the other sources: MRC0222-224 seems to be an intrinsically smaller and more reddened source than the others, while MRC1114-220 may either be an outlier, or have a precessing or misaligned disk. When deprojected using the best-fit disk angles, the source sizes are consistent with a model in which the accretion disk is perpendicular to the radio jets, and the heads of the radio jets expand at approximately constant speed up to a size $\\sim 1$ Mpc; the paucity of sources larger than this limit can be explained by the average duty cycle of these AGN. The distribution of fitted disk angles is consistent with the calculation of expected jet angles for Doppler-boosted jets with Lorentz factors of $\\sim$ 20, if the accretion disk rotation axis and radio jets are assumed to be coincident. The calculated jet angle distribution may match the disk angle distribution better if a more physically reasonable model was used for the jet, with a core of higher Lorentz factor than the surrounding jet sheath, as opposed to this simple single Lorentz factor model. There is a weak correlation ($> 1.5 \\sigma$ significance by Kendall's $\\tau$ test) between the low-frequency radio luminosity and the sine of the angle between the accretion disk axis and the line of sight. This is predicted by the receding torus model of \\citet{lawrence91} if the accretion disks are perpendicular to the radio jets, such that the disk angles match the jet angles. In this model, the dust is sublimated by the AGN nucleus, leading to larger opening angles of the obscuring dusty torus for more radio-luminous sources, and hence quasars are seen up to larger jet angles. The largest accretion disk angle measured is 48$\\degree$, which is consistent with the opening angle predicted for powerful FRII radio sources. A larger sample size is needed to confirm the radio luminosity -- disk angle correlation, but this is vital, as it would yield the first direct test of the receding torus model. It should be noted that when the analysis was performed again with the models in which the accretion disk was fixed at the redshift of the NLR, this result disappeared, and so the radio luminosity -- disk angle correlation is strongly dependent on the model used. A study of the velocity shifts between the line components revealed the single-peaked broad lines to be formed at similar redshifts to the narrow lines. The disk emission tended to be redshifted with respect to these lines for high-$z${} sources and blueshifted for lower-$z$ sources. Since the accretion disk is expected to be at the systemic velocity, this would imply that the single-peaked broad lines and the narrow lines are formed in outflows for more luminous, high-$z${} sources, and in infalling clouds for closer, less luminous sources. Such a scenario could arise if powerful winds were causing outflows of gas from the high-$z$, luminous sources, while these winds have stalled in the less-luminous quasars seen at $z < 1$, resulting in infall of the gas back towards the galaxy nucleus. The results of this paper are all dependent on the validity of the disk model used in the Bayesian fitting. This model, taken from \\citet{chen89}, describes a circular, Keplerian disk. The illuminating flux from the central black hole was modelled as falling off with a radial exponent of 3. In order to investigate the effects of range of permutations to this basic model, such as a variety of disk emissivities and warped or elliptical disks, greater spectral resolution, a larger quasar sample size, and in particular, multi-epoch data to study the variability of the emission line profiles are required." }, "0910/0910.2701_arXiv.txt": { "abstract": "The ``Mouse'' (PWN~G359.23$-$0.82) is a spectacular bow shock pulsar wind nebula, powered by the radio pulsar J1747$-$2958. The pulsar and its nebula are presumed to have a high space velocity, but their proper motions have not been directly measured. Here we present 8.5 GHz interferometric observations of the Mouse nebula with the Very Large Array, spanning a time baseline of 12 yr. We measure eastward proper motion for PWN~G359.23$-$0.82 (and hence indirectly for PSR~J1747$-$2958) of $12.9\\pm1.8$~mas~yr$^{-1}$, which at an assumed distance of 5~kpc corresponds to a transverse space velocity of $306\\pm43$~km~s$^{-1}$. Considering pressure balance at the apex of the bow shock, we calculate an in situ hydrogen number density of approximately $1.0_{-0.2}^{+0.4}$~cm$^{-3}$ for the interstellar medium through which the system is traveling. A lower age limit for PSR~J1747$-$2958 of $163_{-20}^{+28}$~kyr is calculated by considering its potential birth site. The large discrepancy with the pulsar's spin-down age of 25~kyr is possibly explained by surface dipole magnetic field growth on a timescale $\\approx$15~kyr, suggesting possible future evolution of PSR~J1747$-$2958 to a different class of neutron star. We also argue that the adjacent supernova remnant G359.1$-$0.5 is not physically associated with the Mouse system but is rather an unrelated object along the line of sight. ", "introduction": "The evolution of neutron stars and the potential relationships between some of their observed classes remain outstanding problems in astrophysics. Proper motion studies of neutron stars can provide independent age estimates with which to shed light on these questions. In particular, the well defined geometry of bow shock pulsar wind nebulae (PWNe; \\citeauthor{gaensler:3} \\citeyear{gaensler:3}), where the relativistic wind from a high-velocity pulsar is confined by ram pressure, can be used as a probe to aid in the understanding of both neutron star evolution and the properties of the local medium through which these stars travel. The ``Mouse'' (PWN~G359.23$-$0.82), a non-thermal radio nebula, was discovered as part of a radio continuum survey of the Galactic center region \\citep{yusef}, and was suggested to be powered by a young pulsar following X-ray detection \\citep{predehl}. It is now recognized as a bow shock PWN moving supersonically through the interstellar medium (ISM; \\citeauthor{gaensler:5} \\citeyear{gaensler:5}). Its axially symmetric morphology, shown in Figure \\ref{fig:yusef20cm}, consists of a compact ``head'', a fainter ``body'' extending for $\\sim$10$^\\prime$$^\\prime$, and a long ``tail'' that extends westward behind the Mouse for $\\sim$40$^\\prime$$^\\prime$ and $\\sim$12$^\\prime$ at X-ray and radio wavelengths respectively \\citep{gaensler:5,mori}. The cometary tail appears to indicate motion away from a nearby supernova remnant (SNR), G359.1$-$0.5 \\citep{yusef}. \\begin{figure*}[t] \\setlength{\\abovecaptionskip}{-7pt} \\begin{center} \\includegraphics[trim = 0mm 0mm 0mm 0mm, clip, angle=-90, width=13cm]{f1.ps} \\end{center} \\caption{VLA image of the Mouse (PWN~G359.23$-$0.82) at 1.4~GHz with a resolution of 12\\farcs8$\\times$8\\farcs4 (reproduced from \\citeauthor{gaensler:5} \\citeyear{gaensler:5}). The brightness scale is logarithmic, ranging between $-$2.0 and $+$87.6~mJy~beam$^{-1}$ as indicated by the scale bar to the right of the image. The eastern rim of SNR~G359.1$-$0.5 is faintly visible west of $\\sim$RA~17$^{\\mbox{\\scriptsize{h}}}$46$^{\\mbox{\\scriptsize{m}}}$25$^{\\mbox{\\scriptsize{s}}}$.} \\label{fig:yusef20cm} \\end{figure*} A radio pulsar, J1747$-$2958, has been discovered within the ``head'' of the Mouse \\citep{camilo:1}. PSR~J1747$-$2958 has a spin period $P=98.8$ ms and period derivative $\\dot{P}=6.1\\times10^{-14}$, implying a spin-down luminosity $\\dot{E}=2.5 \\times 10^{36}$~ergs~s$^{-1}$, surface dipole magnetic field strength $B=2.5\\times10^{12}$~G, and characteristic age $\\tau_{c} \\equiv P/ 2\\dot{P}=25$~kyr (\\citeauthor{camilo:1} \\citeyear{camilo:1}; see also updated timing data from \\citeauthor{gaensler:5} \\citeyear{gaensler:5}). The distance to the pulsar is {\\footnotesize $\\gtrsim$}4~kpc from X-ray absorption \\citep{gaensler:5}, and {\\footnotesize $\\lesssim$}5.5 kpc~from HI absorption \\citep{uchida}. Here we assume that the system lies at a distance of $d=5d_{5}$~kpc, where $d_{5}=1\\pm0.2$ ($1\\sigma$). Given such a small characteristic age, it is natural to ask where PSR~J1747$-$2958 was born and to try and find an associated SNR. While it is possible that no shell-type SNR is visible, such as with the Crab pulsar \\citep{sankrit} and other young pulsars \\citep{braun}, an association with the adjacent SNR~G359.1$-$0.5 appears plausible. This remnant was initially suggested to be an unrelated background object near the Galactic center \\citep{uchida}. However, it is now believed that the two may be located at roughly the same distance (\\citeauthor{yusef3} \\citeyear{yusef3}, and references therein). By determining a proper motion for PSR~J1747$-$2958, this association can be subjected to further scrutiny (for example, see analysis of PSR~B1757$-$24, PWN~G5.27$-$0.90 and SNR~G5.4$-$1.2; \\citeauthor{blazek} \\citeyear{blazek}; \\citeauthor{zeiger} \\citeyear{zeiger}). As PSR~J1747$-$2958 is a very faint radio source, it is difficult to measure its proper motion interferometrically. It is also difficult to use pulsar timing to measure its proper motion due to timing noise and its location near the ecliptic plane \\citep{camilo:1}. To circumvent these issues, in this paper we investigate dual-epoch high-resolution radio observations of the Mouse nebula, spanning 12 years from 1993 to 2005, with the intention of indirectly inferring the motion of PSR~J1747$-$2958 through the motion of its bow shock PWN. In \\S~\\ref{SectionObservations} we present these observations. In \\S~\\ref{SectionAnalysis} we present our analysis and measurement of proper motion using derivative images of PWN~G359.23$-$0.82. In \\S~\\ref{SectionDiscussion} we use our measurement to determine an in situ hydrogen number density for the local ISM, to resolve the question of a possible association with SNR~G359.1$-$0.5, and to investigate the age and possible future evolution of PSR~J1747$-$2958. We summarize our conclusions in \\S~\\ref{SectionConclusions}. ", "conclusions": "We have investigated two epochs of interferometric data from the VLA spanning 12 years to indirectly infer a proper motion for the radio pulsar J1747$-$2958 through observation of its bow shock PWN~G359.23$-$0.82. Derivative images were used to highlight regions of rapid spatial variation in flux density within the original images, corresponding to the vicinity of the forward termination shock, thereby acting as a proxy for the motion of the pulsar. We measure an eastward proper motion for PWN~G359.23$-$0.82 of $\\mu=12.9\\pm1.8$~mas~yr$^{-1}$, and interpret this value as an upper limit on the motion of PSR~J1747$-$2958. At this angular velocity, we argue that PSR~J1747$-$2958 is moving too slowly to be physically associated with the relatively young adjacent SNR~G359.1$-$0.5, independent of distance estimates to either object or of inclination effects. At a distance $d=5d_{5}$~kpc, the proper motion corresponds to a projected velocity of $V_{\\mbox{\\tiny{PSR,$\\perp$}}}$~{\\footnotesize $\\leq$}~$\\left(306\\pm43\\right)d_{5}$~km~s$^{-1}$, which is consistent with the projected velocity distribution for young pulsars. Combining the time taken for PSR~J1747$-$2958 to traverse its smooth $\\sim$12$^\\prime$ radio tail with the time to escape a typical SNR, we calculate a lower age limit for PSR~J1747$-$2958 of $t_{\\mbox{\\tiny{total}}}$~{\\footnotesize $\\geq$}~$163_{-20}^{+28}$~kyr (68\\% confidence). The lower age limit $t_{\\mbox{\\tiny{total}}}$ exceeds the characteristic age of PSR~J1747$-$2958 by more than a factor of 6, arguably providing the most robust evidence to date that some pulsars may be much older than their characteristic age. This age discrepancy for PSR~J1747$-$2958 suggests that the pulsar's spin rate is slowing with an estimated braking index $n$~{\\footnotesize $\\lesssim$}~$1.3$ and that its magnetic field is growing on a timescale $\\approx$15~kyr. Such potential for magnetic field growth in PSR~J1747$-$2958, in combination with other neutron stars that transcend their archetypal categories such as PSR~J1718$-$3718, a radio pulsar with a magnetar-strength magnetic field that does not exhibit magnetar-like emission \\citep{kaspi}, PSR~J1846$-$0258, a rotation-powered pulsar that exhibits magnetar-like behaviour \\citep{gavril,archibald}, and magnetars such as 1E 1547.0$-$5408 that exhibit radio emission (\\citeauthor{camilo:2} \\citeyear{camilo:2}, and references therein), supports the notion that there may be evolutionary links between the rotation-powered and magnetar classes of neutron stars. However, such a conclusion may be difficult to reconcile with evidence suggesting that magnetars are derived from more massive progenitors than normal pulsars (e.g., \\citeauthor{gaens:8} \\citeyear{gaens:8}; \\citeauthor{muno} \\citeyear{muno}). If the massive progenitor hypothesis is correct, then this raises further questioning of whether, like the magnetars, there is anything special about the progenitor properties of neutron stars such as PSR~J1747$-$2958, or whether all rotation-powered pulsars exhibit similar magnetic field growth or even magnetar-like phases in their lifetimes. To constrain the motion of PSR~J1747$-$2958 further, future observational epochs are desirable. It may be possible to better constrain the motion and distance to this pulsar by interferometric astrometry with the next generation of sensitive radio telescopes (e.g., \\citeauthor{2004NewAR..48.1413C} \\citeyear{2004NewAR..48.1413C}). High time resolution X-ray observations may also be useful to detect any magnetar-like behaviour from this rotation-powered radio pulsar. In general, more neutron star discoveries, as well as measured or inferred braking indices, may allow for a better understanding of possible neutron star evolution." }, "0910/0910.0368_arXiv.txt": { "abstract": "The role of null-point reconnection in a 3D numerical MHD model of solar emerging flux is investigated. The model consists of a twisted magnetic flux tube rising through a stratified convection zone and atmosphere to interact and reconnect with a horizontal overlying magnetic field in the atmosphere. Null points appear as the reconnection begins and persist throughout the rest of the emergence, where they can be found mostly in the model photosphere and transition region, forming two loose clusters on either side of the emerging flux tube. Up to 26 nulls are present at any one time, and tracking in time shows that there is a total of 305 overall, despite the initial simplicity of the magnetic field configuration. We find evidence for the reality of the nulls in terms of their methods of creation and destruction, their balance of signs, their long lifetimes, and their geometrical stability. We then show that due to the low parallel electric fields associated with the nulls, null-point reconnection is not the main type of magnetic reconnection involved in the interaction of the newly emerged flux with the overlying field. However, the large number of nulls implies that the topological structure of the magnetic field must be very complex and the importance of reconnection along separators or separatrix surfaces for flux emergence cannot be ruled out. ", "introduction": "\\label{sec:intro} The continual injection of new magnetic flux, energy, and helicity into the Sun's atmosphere from the convection zone below is fundamentally responsible for much of the activity observed in the solar atmosphere. As the newly emerged flux interacts with the pre-existing magnetic field above, X-ray jets, bright points, or active-region loop systems may be formed (depending on the spatial scale of the emergence), thus these features are a result of magnetic reconnection and so are the realisation of the heated corona. \\subsection{MHD modelling of flux emergence} Numerical simulations are a powerful tool for modelling and understanding such complex inherently 3D systems. There is a large body of literature using magnetohydrodynamic (MHD) codes to model flux emergence events. Magnetic flux is created by the solar dynamo in the convective overshoot layer at the base of the convection zone, known as the tachocline \\citep{1992A&A...265..106S}. Instabilities cause tubes of flux to become buoyant and rise through the convection zone towards the solar surface. During the initial rise phase, the thin-flux-tube approximation has been used to predict active-region emergence latitudes and tilt angles that are consistent with observations \\citep{1994A&A...281L..69S,1995ApJ...441..886C,2006MNRAS.369.1703H}. However, as the flux tubes approach the photosphere they expand significantly, so the thin-flux-tube approximation is no longer valid, as the flux tubes are likely to become fragmented or at least severely distorted, and full 3D MHD models are required. The magnetic fieldlines must have enough twist to prevent the interactions with the surrounding medium from fragmenting the flux tube \\citep{1996ApJ...464..999L}, although the critical degree of twist required depends on the apex curvature of the loop and can be significantly less in 3D than the 2D limit implies \\citep{2000ApJ...540..548A}. The emergence in three dimensions of a single flux tube or sheet into an initially field-free atmosphere was first studied by \\citet{1992PASJ...44..167M}, who obtained many basic features of the emerging field, such as draining of plasma down the fieldlines, expansion of loops into the corona, and formation of shock waves at the loop footpoints. Later work on this topic by several authors, all using a stably stratified convection zone, a low temperature photosphere and a high temperature (but field-free) corona, went into greater detail \\citep[see ][]{2005ApJ...635.1299A}. \\citet{2003ApJ...582..475A}, using an anelastic MHD model of the convection zone coupled to a fully compressible code for the atmosphere, found that the newly-emerged coronal fieldlines formed sigmoidal structures whose chirality depended on where in the atmosphere the emitting plasma was assumed to be located. \\citet{2003ApJ...586..630M} showed that the initially bipolar photospheric magnetic field structure can develop into a quadrupolar structure, and developed the classification of emerged fieldlines as either expanding or undulating. Finally, \\citet{2004ApJ...610..588M} found that the natural shearing motions that occur as the field expands into a pressure-stratified atmosphere tend to add axial flux and decrease the twist of the fieldlines. However, the real corona contains a pre-existing magnetic field, and interesting results have been obtained by including an overlying field in flux emergence models. In 2.5D, \\citet{1996PASJ...48..353Y} experimented with vertical, oblique, and several orientations of horizontal background coronal magnetic fields. All the configurations lead to a current sheet forming over the emerging loops, and the production of magnetic islands and jets via the tearing instability. However, as \\citet{2003ApJ...586..630M} have pointed out, full 3D models are required to adequately reproduce the draining of plasma down the magnetic fieldlines. Such simulations have been carried out by, for example, \\citet{2004A&A...426.1047A,2005ApJ...635.1299A}, who allowed a twisted flux tube to emerge into a uniform horizontal background magnetic field. They found that in such a 3D setup, magnetic reconnection occurred at multiple sites along fieldlines, rather than at a single isolated magnetic null as is the norm in 2D. This model reproduced fan-like arcades similar to those observed by TRACE, and also reconnection jets whose temperatures and velocities were a good fit with observations. \\citet{2004ApJ...609.1123F} also emerged a 3D twisted magnetic flux tube into a pre-existing coronal field, although here the coronal field was a potential arcade. Their model does not include a convection zone; the flux tube is `injected' into the atmosphere by controlling the value of the electric field ($\\mathbf{v} \\times \\mathbf{B}$) at the lower boundary, thus specifying the electric current and hence the Lorentz force. This allows them to achieve full emergence of the tube into the atmosphere, by analytically defining the shape of the field, and the velocity with which it passes through the boundary. Although the ideal MHD equations were used, magnetic reconnection could still take place at current sheets due to numerical diffusion in the presence of strong gradients in the magnetic field. Once a certain critical amount of twist had emerged, the flux tube underwent the kink instability and began to accelerate upwards and twist around, producing a sigmoidal current sheet beneath. This evolution was found to be strongly dependent on the relative orientations of the emerging flux tube and the pre-existing coronal magnetic field: if the overlying field was oppositely-directed to the axial component of the flux tube, reconnection started straightaway and destroyed the flux tube as it rose into the atmosphere. The effect of the orientation of the overlying field was investigated by \\citet{2007ApJ...666..516G}, using the same model as the Archontis papers above, but with different orientations of the plane-parallel coronal field relative to the emerging flux tube. They found that most reconnection took place when the two flux systems were close to anti-parallel, and hardly any when they were close to parallel, but, in all cases, the height reached by the emerging flux tube, as a function of time, was unchanged. The models described so far all made certain simplifications about the behaviour of the convection zone, since their main aim was to investigate the effects of emergence into the solar atmosphere. However, the most realistic models currently possible of magnetic flux rising through the convection zone include radiative transfer, convection, compressibility, and thermal conduction. For example, \\citet{2007A&A...467..703C} showed how a single initial rising flux tube influenced by convection leads to the emergence of many small flux bundles, which produce a characteristic (observed) dark lane appearing in the photospheric granulation pattern. However, since they have no atmosphere above the photosphere they could not study the effects of emergence on the atmosphere itself. \\citet{2008ApJ...679..871M} recently performed a simulation in which their corona was heated by the release of stress built up by convective motions in a complex magnetic field topology. The granulation cells became enlarged as an emerging flux tube passed through them, in agreement with observations. Recent work by \\citet{2007A&A...470..709M} has shown that in fact such a high level of complexity in the subsurface field is not necessary for emergence models to give valid results in the solar atmosphere. They create a complex subsurface field by allowing two magnetic flux tubes to interact in a stably stratified convection zone before emerging, and show that the resulting emerged atmospheric magnetic field is very similar to the result for a simple single flux tube. This means that the results obtained from simpler models with stably-stratified convection zones are adequate when the aim is to study the resulting atmospheric magnetic fields. \\subsection{The next step: magnetic topology} As we have just seen, previous work on solar emerging flux has greatly improved our understanding of many of the complex processes involved. But there is one key issue that was not addressed in all the previous analyses; they do not account for the topology of the magnetic fields. Here, by magnetic topology, we mean specifically the location and evolution of topological features such as magnetic null points, spines, separatrix surfaces, and separators \\citep[see][for a good review]{2005LRSP....2....7L}. Plotting individual fieldlines that are not part of the topological skeleton, even if they are carefully selected, is not enough to ensure that all the information about the structure of the magnetic field, and the way it connects, is known. Magnetic reconnection preferentially occurs at topological features including null points \\citep{2004GApFD..98..407P,2005GApFD..99...77P,2007JGRA..11203103P}, separatrix surfaces \\citep{2002ApJ...576..533P,2005ApJ...624.1057P}, and separator fieldlines \\citep{2005ApJ...624.1057P,2007RSPSA.463.1097H,2008ApJ...675.1656P}, as well as at their geometrical counterparts: quasi-separatrix layers \\citep{2006SoPh..238..347A,2006AdSpR..37.1269D,2007Sci...318.1588A,2007ApJ...660..863T} and quasi-separators \\citep[also known as hyperbolic flux tubes;][]{2003ApJ...582.1172T,2003ApJ...595..506G,2005A&A...444..961A,2006A&A...451.1101D,2006A&A...459..627D,2007A&A...473..615W}. This is because the hyperbolic magnetic topologies around such features tend to focus the electric current, and a strong electric field (parallel to the magnetic field) is associated with 3D magnetic reconnection. So knowing the topological structure of the field provides a good guide to the probable locations of reconnection sites in the system. In this paper, we concentrate on determining the importance of reconnection at magnetic null points in our model magnetic field. The signature of 3D magnetic reconnection is the existence of a strong electric field parallel to the magnetic field \\citep{1988JGR....93.5547S}. We therefore study the parallel electric field near the nulls, to determine if reconnection takes place there. We also consider the current density, which may be parallel ($\\mathbf{j}_{\\parallel}$) or perpendicular ($\\mathbf{j}_{\\perp}$) to the spine of the null point, or a combination of both, and the consequences of this for the structure of the surrounding fieldlines are discussed in Section~\\ref{ssec:locnat}. 3D magnetic reconnection may take place in the presence or absence of magnetic null points \\citep{1988JGR....93.5547S,1988JGR....93.5559H,1995JGR...10023443P,1996JGR...101.7631D,1996RSPSA.354.2951P,2003PhPl...10.2717H,2006RSPSA.462.2877W}. However, in the flux emergence model that we study here, the change of magnetic topological structure is a central feature, and is cospatial with the location of magnetic reconnection. To have a changing magnetic topology, nulls are often involved, and hence we focus on reconnection at nulls in this paper. Reconnection at other topological features in our flux emergence experiment may also be significant, and this will form the subject matter for a future investigation. So, the magnetic topology is important as it allows us to identify probable sites of magnetic reconnection. However, it has only been realised very recently that information about the topological structure is also required to make sense of the reconnection rate and energy release sites. Knowing the rate at which magnetic reconnection takes place is crucial to our understanding of how the magnetic field evolves, and recent work by \\citet{2008ApJ...675.1656P} has shown that knowledge of the magnetic topology is absolutely essential for a correct interpretation of the results of the model. This is because, in a complicated magnetic field containing fieldlines with many different connectivities, a phenomenon called \\emph{recursive reconnection} can take place. Recursive reconnection means that the same magnetic flux can be recycled many times through a repeating sequence of different magnetic connectivities. In a situation like this, knowledge of the topological structure of the magnetic field is vital to determine the global reconnection rate, which can in fact be much higher than would be determined via other methods. In this paper, we take the first steps towards a full topological analysis of a flux emergence model. The dataset that we use is one of the numerical MHD models of \\citet{2007ApJ...666..516G}, which consists of a twisted buoyant magnetic flux tube rising through the upper layers of a stably-stratified convection zone and into an atmosphere with a horizontal plane-parallel pre-existing magnetic field. Now, all of the fieldlines in the topological skeleton of the magnetic field must either start or end at magnetic null points (there are no bald patches \\citep{1993A&A...276..564T} in our model as no fieldlines leave the closed boundaries of our box, only the periodic boundaries). We locate the null points in the magnetic field using the accurate new algorithm described by \\citet{2007PhPl...14h2107H}. This paper concentrates on the surprising nature of these magnetic null points: their number, type, distribution, and evolution. Section~\\ref{sec:model} describes the code, the model setup, and how it was analysed. Our results are reported in Section~\\ref{sec:results} and then we conclude with a discussion in Section~\\ref{sec:conclusions}. ", "conclusions": "\\label{sec:conclusions} In this paper, we have analysed the properties of the magnetic null points present in a model solar flux emergence event, and considered the implications for magnetic reconnection. The flux emergence model that we used was first described by \\citet{2007ApJ...666..516G}. In it, a twisted magnetic flux tube rises through a stratified solar convection zone and atmosphere to interact and reconnect with a horizontal overlying magnetic field. We detected a large number of magnetic null points within the resulting model magnetic field. Nulls begin to appear when the magnetic reconnection starts, and persist until the end of the entire model run. Up to 26 nulls are present at any one time. The nulls were tracked through all the time-frames of the model, and we found that they number 305 in total. Our surprise at this unexpectedly-large number of nulls led us to speculate that some of them might simply be numerical artefacts. However, there is a lot of evidence to show that the vast majority of nulls are real, are not due to numerical artefacts, and that we can reliably identify and exclude those that are. The nulls were also classified by sign (positive or negative), and we found that the balance of signs is always in agreement with a conservation law derived from the 3D Euler equation (equation~\\ref{eq:euler}), once the spurious nulls are excluded. The lifetimes of the nulls were calculated, and the majority of the nulls were found to persist for many multiples of the mean timestep between frames. Finally, the stability of the nulls was also assessed by calculating how fast their spine eigenvectors were rotating. The vast majority of nulls were found to have remarkably steady spine directions, meaning that they are very stable. The nulls are located throughout the model photosphere and transition region. They never reach as high as the bottom of the model corona. The reason for this is because in the vast majority of our model corona the magnetic field is very simple (an overlying horizontal magnetic field) and the emergence of new flux only creates complexity in small localised regions low down in the atmosphere. \\citet{2009SoPh..254...51L} have investigated the number of nulls that occur in the solar atmosphere and showed that indeed their number falls off quickly the further you get from the photosphere, as the complexity of the magnetic field decreases with height. The nulls in our experiment form just after the reconnection has started, in two clumps low down at the boundary between the emerging and the overlying magnetic flux. As the emerging region expands, the nulls are pushed out towards the side boundaries of the box, but they never reach or cross these boundaries during the model run. A parallel electric field is a necessary requirement for reconnection, so we investigated its nature in the vicinity of the nulls. Regions of strong parallel electric field are found in the model at the peak of the emerging fieldlines, and also (later on) lower down in the atmosphere. Although many of the nulls have some associated parallel electric field, we found that this is weak compared to the strongest parallel electric fields elsewhere in the atmosphere, which are concentrated far from the nulls in the regions just described. Null-point reconnection is therefore ruled out as the predominant type of magnetic reconnection taking place during the emergence event. Clearly though, the reconnection is associated with either separators from these nulls or reconnection in the absence of nulls. We plan to investigate these possibilities more thoroughly in future. \\begin{acks} The authors would like to thank Dr.\\ D.\\ Pontin for his insightful comments, and also the anonymous referee for his thoughtful suggestions which helped us to improve the paper. Computational time on the UKMHD Linux clusters in St Andrews (STFC and SRIF funded) is gratefully acknowledged. This work was supported by the European Commission through the SOLAIRE Network. RCM is financially supported by the University of St Andrews Solar Group STFC Rolling Grant. \\end{acks}" }, "0910/0910.3513.txt": { "abstract": "{ Accurate photometric CoRoT space observations of a secondary seismological target, \\hk, led to the discovery that this star is an astrophysically important double-lined eclipsing spectroscopic binary in an eccentric orbit ($e\\sim0.3$), unusual for its short 3\\fd65705 orbital period. The high eccentricity, coupled with the orientation of the binary orbit in space, explains the very unusual observed light curve with strongly unequal primary and secondary eclipses having the depth ratio of 1-to-100 in the CoRoT ``seismo\" passband. Without the high accuracy of the CoRoT photometry, the secondary eclipse, 1.5 mmag deep, would have gone unnoticed. A spectroscopic follow-up program provided 45 high dispersion spectra. The analysis of the CoRoT light curve was performed with an adapted version of PHOEBE that supports CoRoT passbands. The final solution was obtained by simultaneous fitting of the light and the radial velocity curves. Individual star spectra were derived by spectrum disentangling. The uncertainties of the fit were derived by bootstrap resampling and the solution uniqueness was tested by heuristic scanning. The results provide a consistent picture of the system composed of two late B stars. The Fourier analysis of the light curve fit residuals yields two components, with orbital frequency multiples and an amplitude of $\\sim$ 0.1 mmag, which are tentatively interpreted as tidally induced pulsations. An extensive comparison with theoretical models is carried out by means of the Levenberg-Marquardt minimization technique and the discrepancy between models and the derived parameters is discussed. The best fitting models yield a young system age of 125 million years which is consistent with the eccentric orbit and synchronous component rotation at periastron.\\\\} ", "introduction": "\\object{HD 174884} (Corot 7758) was selected as a seismology target for the first short run in the CoRoT ``center\" direction (SRc1) because -- on the basis of its B8~V spectral classification -- it was considered a good candidate to pinpoint the instability strip red edge of the Slowly Pulsating B stars \\citep[e.g.][]{wae98}. Its only available spectrum, obtained before launch with the Elodie spectrograph and available from the Gaudi database\\footnote{GAUDI (sdc.laeff.inta.es/gaudi/) is the data archive of the ground-based asteroseismology programme of the CoRoT mission. The GAUDI system is maintained at LAEFF, which is part of the Space Science Division of INTA.} \\citep{sol05} did not reveal the star's binary nature, which was instead immediately evident from CoRoT photometry. The system was continuously observed during 27 days. The light curve of HD~174884 is characterized by signatures of high eccentricity: the secondary minimum markedly displaced from phase 0.5 and a bump, typical of close eccentric binaries, due to variable distortion of stellar surfaces along the orbit. The system is very interesting for several reasons: the high eccentricity is coupled with quite short a period ($P_{orb}=3 \\fd 66$), putting the system in an extreme location of the period - eccentricity diagram of B stars. This indicates young age or an inefficient circularization process. Secondly, the huge difference between the primary and secondary minima, the latter detected only thanks to CoRoT accuracy, and the overall shape and accuracy of the light curve is a challenge for light curve fitting methods and a test of their adequacy to solve light curves of this unprecedented quality. Finally, only a thorough analysis of HD~174884 could yield information of the possible pulsational properties of the components. \\begin{figure*}[!ht] \\centering \\includegraphics[width=17.4cm]{13311fg1.eps} % compressed version \\caption{Upper panel: the complete CoRoT light curve of HD 174884 as normalized flux vs HJD. Lower panel: a blow-up, spanning approximately three periods, showing the tiny secondary minimum (first occurrence at reduced HJD of around 4218) } \\label{lc} \\end{figure*} In the following sections we present the results of the analysis of CoRoT photometry and of ground-based high dispersion spectroscopy. A critical analysis of parameter uncertainties and of the uniqueness of the light and radial velocity curve solution is an essential part of the study. The well constrained physical parameters of the binary components are then compared with evolutionary models computed with the CL\\'ES code \\citep[Code Li\\`{e}geois d' \\'{E}volution Stellaire,][]{scuetal08} with the aim of constraining the age and the structural parameters (chemical composition, overshooting). Finally the light curve fit residuals, with variations of a few $10^{-4}$ mag, are analyzed. %__________________________________________________________________ ", "conclusions": "\\label{conclusions} This work has been carried out with two aims in mind: to achieve a thorough description of an interesting binary and to test the performance of current binary star modeling on data of unprecedented accuracy, as those available for the \\corot{} seismo-field targets. \\hd lends itself to both purposes: its peculiar light curve suggests an unusual system configuration, worth a detailed study, and it is as well quite stable over the seven observed orbital periods so that, if possible systematic errors arising from binary modeling were present, they would not be hidden by transient intrinsic phenomena. (The latter is indeed the case of the other fully analyzed binary seismo-target of \\corot\\ first runs, \\object{AU Mon} \\citep{desmetal09} whose study reveals the presence of variable circumbinary material). Given the presence of a grazing eclipse, we were concerned about solution uniqueness, a problem which is sometimes overlooked, and we put a great effort in getting a sound estimate of parameter uncertainties, essential for sensible comparison with evolutionary stellar models. We can summarize our main results as follows: \\begin{itemize} \\item{due to the high accuracy of CoRoT data, we were able to unambiguously derive parameters, such as inclination and component sizes, which would otherwise be poorly constrained in a system with grazing eclipses. This fact, the excellent agreement between the results of the different methods applied for the analysis of the spectra, and the extensive check of solution uniqueness allowed a sound estimation of the physical parameters of \\hd and their uncertainties.} \\item{The comparison with stellar models is quite satisfactory but for the temperature difference between the components, which, for the best fitting models, is typically a few hundred degrees higher than that from the light curve solution. We interpreted that as due to the comparison with non rotating stellar models. We could as well improve the agreement with models introducing differences in the physical processes acting in the components, such as overshooting, but Occam's razor arguments favor the simpler, though not exhaustive, explanation. Increasing the free parameters yields, in fact, only a marginal improvement. The comparison with theoretical models provided as well an estimation of the system age and some indication on the component chemical composition.} \\item{The dynamical properties of \\hd are in very good agreement with the predictions of Zahn's theory of circularization and spin-orbit synchronization. The high value of the eccentricity for the period suggests that resonance locking has not been at work in the system.} \\item{A few frequencies are clearly detected in the residuals of the light curve fit at multiples of the orbital frequency. We tentatively interpret them as resonantly excited pulsations. This hypothesis is strengthened by the fact that many g-modes of high radial order and degree $\\ell=2$ exist in the frequency range of tidal frequencies. Among the very few detections of tidally induced pulsations in binaries, \\hd is the only case with two well characterized early type MS components.} \\end{itemize} Finally, it is evident from our results that when dealing with observation accuracy of a few tenths of mmag we are close to the level at which systematics from the model has a relevant impact on the light curve solution. This indicates that the models shall be updated, especially in view of the still higher accuracy curves which will be obtained by current and future space missions such as Kepler \\citep{kepler}. Similarly, the physical properties derived by data of such accuracy should be compared with theoretical models including stellar rotation. %__________________________________________________________________" }, "0910/0910.1587_arXiv.txt": { "abstract": "Supermassive black holes (SMBHs) found in the centers of many galaxies are understood to play a fundamental, active role in the cosmological structure formation process. In hierarchical formation scenarios, SMBHs are expected to form binaries following the merger of their host galaxies. If these binaries do not coalesce before the merger with a third galaxy, the formation of a black hole triple system is possible. Numerical simulations of the dynamics of triples within galaxy cores exhibit phases of very high eccentricity (as high as $e \\sim 0.99$). During these phases, intense bursts of gravitational radiation can be emitted at orbital periapsis, which produces a gravitational wave signal at frequencies substantially higher than the orbital frequency. The likelihood of detection of these bursts with pulsar timing and the Laser Interferometer Space Antenna ({\\it LISA}) is estimated using several population models of SMBHs with masses $\\gtrsim 10^7~{\\rm M_\\odot}$. Assuming 10\\% or more of binaries are in triple systems, we find that up to a few dozen of these bursts will produce residuals $>1$ ns, within the sensitivity range of forthcoming pulsar timing arrays (PTAs). However, most of such bursts will be washed out in the underlying confusion noise produced by all the other 'standard' SMBH binaries emitting in the same frequency window. A detailed data analysis study would be required to assess resolvability of such sources. Implementing a basic resolvability criterion, we find that the chance of catching a resolvable burst at a one nanosecond precision level is $2-50$\\%, depending on the adopted SMBH evolution model. On the other hand, the probability of detecting bursts produced by massive binaries (masses $\\gtrsim 10^7\\msun$) with {\\it LISA} is negligible. ", "introduction": "\\label{intro} It is well established that most galaxies host supermassive black holes (SMBHs) in their centers \\citep{rich98}. In the past decade, compelling evidence of the correlation between the mass of the central SMBH and the bulge velocity dispersion and luminosity has been collected ~\\citep{ferrarese00,gebhardt00,merritt01,tremaine02}, indicating a coevolutionary scenario for SMBHs and their hosts. On a cosmological scale, galaxy formation and evolution can be understood by semi-analytic modeling, where properties of the baryonic matter are followed in the evolving dark matter halos obtained from large-scale models of hierarchical gravitational structure formation. A simple model of galaxy and central SMBH evolution in which every merger of galaxies leads quickly to coalescence of their central black holes can quantitatively reproduce both the SMBH mass-bulge luminosity relation~\\citep{kauffmann00} and the SMBH mass-velocity dispersion relation~\\citep{haehnelt00}. In this general picture, if both of the galaxies involved in a merger host a SMBH, then the formation of a SMBH binary is an inevitable stage of the merging process. Following the merger, the two black holes sink to the center of the merger remnant because of dynamical friction ~\\citep{begelman80}. When the mass (either in gas or stars) enclosed in their orbit is of the order of their own mass, they start to feel the gravitational pull of each other, forming a bound binary. The subsequent binary evolution is, however, still unclear. In order to coalesce, the binary must shed its binding energy and angular momentum; a dynamical process known in literature as `hardening'. A crucial point in assessing the fate of the binary is the efficiency with which it transfers energy and angular momentum to the surrounding gas and stars. The case of SMBH binaries in stellar environments has received a lot of attention in the last decade. The system is usually modeled as a massive binary embedded in a stellar background with a given phase space distribution. The region of phase space containing stars that can interact with the SMBH binary in one orbital period is known as the loss cone~\\citep{frankr76,AS01,milos03}. As the binary evolves, it ejects stars on intersecting orbits via the so called `slingshot mechanism', causing a progressive emptying of the loss cone, which ultimately increases the hardening time scale. Without an efficient physical mechanism for repopulating the loss cone, the binary will never proceed to small separations where coalescence induced by gravitational radiation takes place within a Hubble time. This is known as the stalling or `last parsec' problem \\citep{milos01}. In the last decade, several solutions to the stalling issue have been proposed. Axisymmetric or triaxial stellar distributions may significantly shorten the coalescence timescale \\citep{yu02,merritt04,bercziketal06}. This is bacause the presence of deviations from spherical symmetry can produce ``boxy'' orbits, as seen by \\cite{bercziketal06}. These orbits produce centrophilic stellar orbits and, therefore, replenish the loss-cone. However, more recent calculations by \\cite{AmaroSeoaneSantamaria09} of the outcome of the merger of two clusters initially in parabolic orbits \\citep{ASF06} have not been able to reproduce the rotation necessary to create the unstable bar structure. Other studies have invoked eccentricities of the binary to refill the loss cone, since this effect could alter the cross section for super-elastic scatterings (thus altering the state of the loss cone) and shorten the gap to the onset of gravitational radiation effects (e.g.:~\\citealt{hemsendorf02, aarseth03, bercziketal06,ASF06,ASMF09}). The presence of massive perturbers may also help replenishing the loss cone, boosting the binary hardening rate \\citep{perets07}. On the other hand, in smooth particle hydrodynamics simulations of SMBH binaries in gas-rich environments, efficient hardening induced by the tidal interaction between the binary and the gas medium has been observed, indicating a possible quick coalescence \\citep{escala05,dotti06}. However, current simulations do not have the resolution to follow the binary fate down to the gravitational wave (GW) emission regime, and robust conclusions about its late inspiral and coalescence can not be drawn. In any case, very massive low redshift systems, which are the major focus of our study, are more likely to reside in massive gas poor galaxies and their dynamics is probably dominated by stellar interactions. When scaled to very massive binaries (masses $>10^8\\msun$), the inferred coalescence timescales in a stellar dominated environment are of the order of few Gyrs, indicating that SMBH binaries may be relatively long living systems. If the typical timescale between two subsequent mergers is comparable the SMBH binary lifetime, then a third black hole may reach the nucleus when the binary is still in place, and the formation of SMBH triplets might be a common step in the galaxy formation process. Recent studies of galaxy pairs lead to the conclusion that $30-70$\\% of present day massive galaxies have undergone a major merger since redshift one \\citep{bell06,lin08}, where 'major' means with baryonic mass ratio of the two components larger than $1/3$ or $1/4$ (depending on the study), which is a quite conservative threshold. This means that, on average, all massive galaxies have experienced a merger event in the last ten billion years. Assuming uncorrelated events, and a typical binary lifetime of one billion years, then 10\\% of SMBH binaries may form a triplet. With increasing redshift (and decreasing masses), dynamical timescales become shorter and shorter, implying that triplets may have been more common in the high redshift Universe. In this paper we focus on SMBH triplets, studying their dynamical evolution, GW emission, and detectability. Employing sophisticated three body scattering experiments calibrated on direct-summation {\\sc Nbody} simulations, we study the dynamical evolution of the system, finding surprisingly high eccentricities of the inner SMBH binary (up to $e>0.99$). Even though the triple interaction would possibly lead to an ejection of one or even all SMBHs \\citep{valtonen94}, most of the systems are long living ($\\sim10^9$ yrs, \\cite{HoffmanLoeb07}), and final coalescence is more common than ejection, confirming analytical results by \\cite{makino94}. We model at the leading quadrupole order \\citep{peters63} the bursts of gravitational radiation emitted in the highly eccentric phase, assessing detectability with future GW experiments. Adopting cosmologically and astrophysically motivated models for SMBH formation and evolution, we estimate reliable event rates. In order to cover the low frequencies generated by the expected cosmological population of coalescing SMBH binaries \\citep[e.g.,][]{wl03,ses04,ses05,ses07} or plunges of compact objects such as stellar black holes on to supermassive ones \\citep[see e.g.][for a review and references therein]{Amaro-SeoaneEtAl07}, the space-born observatory {{\\it LISA}} \\citep{bender98} has been planned to be covering the range of frequencies of $\\sim10^{-4}-10^{-1}~{\\rm Hz}$. Moving to even lower frequencies, the Parkes Pulsar Timing Array \\citep[PPTA,~][]{manchester06,manchester07}, the European Pulsar Timing Array \\citep[EPTA,~][]{EPTA} and the North American Nanohertz Observatory for Gravitational Waves \\citep[NANOGrav,~][]{NANOGrav} are already collecting data and improving their sensitivity in the frequency range of $\\sim10^{-8}-10^{-6}$ Hz, and in the next decade the planned Square Kilometer Array \\citep[SKA,~][]{laz09} will provide a major leap in sensitivity. Throughout this paper we consider only very massive systems, with total mass $\\sim 10^8\\msun$. Our goal is to investigate if the high frequency nature of eccentric bursts can provide information about systems which would otherwise emit outside the frequency windows of the planned GW experiments quoted above, by shifting wide (separation $\\gtrsim0.1$ pc) SMBH binaries into the PTA window or by boosting relatively massive (masses $>10^7\\msun$) systems into the {\\it LISA} domain. We note that the bursts analyzed here are different from the `bursts with memory', which arise during the actual coalescence of SMBH binaries and are discussed in~\\citet{pshirkov09} and~\\citet{vanhaasteren09}. The structure of the paper is as follows. In Section \\ref{astro}, we describe our comprehensive study of the dynamics of triple systems and investigate the eccentricity evolution of the inner binary by using direct-summation $N-$body techniques and a statistical 3-body sample calibrated on the $N-$body results. In Section \\ref{gravwaves}, we model the GW signal produced by eccentric bursts and we introduce observable quantities for PTAs and {{\\it LISA}}. In Section 4 we construct detailed populations of emitting SMBH binaries and triplets, and we discuss our results in terms of signal observability and detection rates in Section 5. Lastly, we briefly summarize our results in Section 6. ", "conclusions": "\\label{conclusions} We have addressed in this work three different points in the evolution of triplets of SMBHs in the Universe: The Astrodynamics of the system, the potential GW signature and the detectability. We have performed eight different direct-summation $N-$body simulations, one including more than half a million of particles, to calibrate 1,000 3-body scattering experiments, which include post-Newtonian corrections, in order to have a statistical description of the system. Both numerical tools agree that the inner binary of SMBHs will go through a phase of extremely high eccentricity, which is the motivation for the rest of the work. These three-body excitations of episodic high eccentricity configurations of the close SMBH binary produce interesting GW bursts that may be detectable with forthcoming experiments such as PTAs and {{\\it LISA}}. The extreme eccentricities of such bursts on one hand would leave a very distinctive signature, but on the other require the development of appropriate analysis techniques. To compute likely event rates, we extracted catalogues of merging galaxies from the Millennium Run, and we populated them with SMBHs following the known MBH-bulges relations. We then estimated the fractions of triplets and their eccentricity distribution and we computed the induced signals in both PTAs and the {{\\it LISA}} detector. We found that, depending on the details of the SMBH population model, if the fraction of triplets is $\\ge 0.1$, few to a hundred of GW bursts would be produced at a $>1$ ns level in the PTA frequency domain. Most of the signals will be washed out in the confusion noise due to the emission of `ordinary' low eccentric binaries. However, their peculiar features may guide the development of targeted data analysis techniques, that may help to recognize them even if overwhelmed by the confusion noise. Employing a minimal criterion for source resolvability (which provides a strict lower limit), we found that less than one system may be actually pinned down at ns precisions. By running several dozens of Monte Carlo realization of the signal from the cosmological population of SMBH binaries and triplets we quantified a statistical $2-50$\\% chance of having a resolvable burst in the Universe (assuming 10 yrs of observation). The probability for detection with {\\rm {\\it LISA}} is essentially nil. However, we stress the fact that we focused on systems with ${\\cal M}>10^7\\msun$; our results then simply imply that it is extremely unlikely that a system which would normally emit outside the {\\it LISA} range will produce a burst in the {\\it LISA} window because of resonant three body interactions. On the other hand, if a consistent fraction of {light} binaries (${\\cal M}<10^7\\msun$) is involved in triple systems, we may expect several eccentricity-driven coalescences to be observed by {\\it LISA}. This eventuality would call for the development of extremely eccentric templates ($e>0.9$) for merging SMBHs, and of adequate analysis techniques to extract the signal." }, "0910/0910.4194_arXiv.txt": { "abstract": "We present \\jb, \\hb\\ and \\kb\\ photometry of the Orion Nebula Cluster obtained at the CTIO/Blanco 4~m telescope in Cerro Tololo with the ISPI imager. From the observations we have assembled a catalog of about $\\sim7800$ sources distributed over an area of approximately $30\\arcmin\\times40\\arcmin$, the largest of any survey deeper than 2MASS in this region. The catalog provides absolute coordinates accurate to about 0.15 arcseconds and $3\\sigma$ photometry in the 2MASS system down to $\\jb\\simeq 19.5$~mag, $\\hb\\simeq18.0$~mag, $\\kb \\simeq18.5$~mag, enough to detect planetary size objects 1~Myr old under $A_V\\simeq10$~mag of extinction at the distance of the Orion Nebula. We present a preliminary analysis of the catalog, done comparing the (\\jb-\\hb,\\hb-\\kb) color-color diagram, the (\\hb,\\jb-\\hb) and (\\kb,\\hb-\\kb) color-magnitude diagrams and the \\jb\\hb\\kb\\ luminosity functions of three regions at increasing projected distance from the Trapezium. Sources in the inner region typically show IR\\ colors compatible with reddened T\\ Tauri stars, whereas the outer fields are dominated by field stars seen through an amount of extinction which decreases with the distance from the center. The color-magnitude diagrams make it possible to clearly distinguish between the main ONC population, spread across the full field, and background sources. % The luminosity functions of the inner region, corrected for completeness, remain relatively flat in the sub-stellar regime regardless of the strategy adopted to remove background contamination. ", "introduction": "\\label{sec:intro} The Orion Nebula hosts the richest cluster of\\ young ($\\tau\\simeq 1$~Myr) Pre-Main-Sequence (PMS) stars within 1~kpc of the Sun and therefore represents an ideal laboratory to understand the process of star formation \\citep[see][for recent reviews]{Muench+08, ODell+08}. The cluster, both rich ($n\\approx 2000$ members) and dense (about $2\\times10^4$\\ sources per cubic parsec at its center), is dominated by a small number of massive OB\\ stars mostly clustered in the Trapezium multiplet ($\\theta^1$~Ori). Their UV\\ emission has created a blister HII\\ \\ region ({\\sl Orion Nebula, M~42, NGC~1976}) whose ionization front is still carving the underlying Orion Molecular Cloud \\cite[OMC-1,][] {ODell+08,ODell+09}. About half of the young cluster members, surrounded by their original circumstellar disks, have been already exposed to the hard-UV\\ radiation of the Trapezium stars, whereas the other half remain enshrouded within the OMC-1, together with newer active sites of star formation. Visible and near-IR\\ data, both in imaging and spectroscopy, are needed to characterize the main physical parameters of the cluster population \\citep{LAH97}. These observations, however, are hampered by the brightness and non-uniformity of the nebular background. To overcome these difficulties, we have exploited the unique combination of sensitivity and spatial resolution offered by the Hubble Space Telescope (HST)\\ to obtain accurate photometry of the cluster, especially at substellar masses (HST\\ Treasury program GO-10246). Our extensive HST\\ survey has been complemented by ground-based observations, imaging the Orion Nebula from the {\\sl U}-band to the \\kb-band at La Silla and Cerro Tololo (CTIO)\\ observatories. The ground-based observations, carried out in parallel on the same nights (but at a different epoch than the HST observations), complement the deep HST data which saturate at relatively low brightness levels. Their simultaneity makes the derived stellar colors largely immune to the uncertainties associated with source variability. The ground-based survey at visible wavelengths, performed with the Wide Field Camera (WFI) at the ESO/MPG 2.2~m telescope at La Silla, has been recently presented in \\cite{DaRio+09}. In this paper we present the ground-based IR\\ survey, performed at the CTIO/Blanco telescope with the ISPI\\ imager. Its combination of sensitivity and field coverage (about $30\\arcmin\\times40\\arcmin$) ideally complements the previous surveys of this region. In particular, it reaches fainter magnitudes than the wide field survey ($30\\arcmin\\times45\\arcmin$) by Hillenbrand et al. (1998), while it covers a much larger area than the deep surveys done with Keck (\\hb\\kb-bands, $5\\farcm1 \\times 5\\farcm1$, Hillenbrand \\&\\ Carpenter, 2000), HST-NICMOS\\ (\\jb\\hb-bands, $2\\farcm3\\times2\\farcm3$, Luhman et al. 2000), UKIRT\\ ({\\sl I}\\jb\\hb-bands,\\ 36 arcmin$^2$, Lucas \\&\\ Roche 2000), NTT\\ (\\jb\\hb\\kb, $5\\arcmin\\times5\\arcmin$, Muench et al. 2002) and Gemini (\\jb\\hb\\kb-bands, 26~arcmin$^2$, Lucas et al. 2005). All these deep surveys concentrate on a field centered around the Trapezium stars, or in its immediate surroundings in the case of \\cite{LR00}. In Section\\ \\ref{sec:obs} we discuss the observing strategy, while in Section\\ \\ref{sec:reduction} we describe the data reduction and photometric calibration. In Section\\ 4 we present our completeness analysis. The resulting source catalog is presented in Section 5, whereas in Section 6 we discuss the color-color and color-magnitude diagrams and the luminosity function derived from our \\jb, \\hb, and \\kb-band photometry. ", "conclusions": "We have presented a photometric survey of the Orion Nebula Cluster in the \\jb-, \\hb- and \\kb-passbands carried out at the 4m telescope of Cerro Tololo. The survey, covering a field of about $30\\arcmin\\times40\\arcmin$ centered about $1\\arcmin$\\ southwest of the Trapezium, has been performed in parallel to visible observations of the same region made in La Silla (Da Rio et al. 2009). The two datasets constitute the first panchromatic survey covering simultaneously the Orion Nebula Cluster from the {\\sl U}-band to the \\kb-band. \\ The final catalog, photometrically and astrometrically calibrated to the 2MASS\\ system, contains 7759 sources, (including 174 and 22 sources whose photometry has been taken directly from the 2MASS\\ and \\cite{Mue02} catalogs, respectively). This represents the largest near-IR\\ catalog of the Orion Nebula Cluster to date. Our sensitivity limits allow to detect objects of a few Jupiter masses under about $A_V\\simeq 10$ magnitudes of extinction. We present the color-color diagrams, color-magnitude diagrams and the luminosity functions for three regions centered on the Trapezium containing the same number of sources (excluding the M43 sub-cluster). Sources in the inner region typically show IR\\ colors compatible with reddened T\\ Tauri stars, whereas the outer fields are dominated by field stars seen through an amount of extinction which decreases with the projected distance from the center. The color-magnitude diagrams allow to clearly distinguish between the main ONC population, spread across the full field, and background sources. After correction for completeness, the luminosity functions in the inner region remains nearly flat, with marginal contamination from background sources." }, "0910/0910.1913_arXiv.txt": { "abstract": "We present optical photometry and spectra for the Type Ia supernova (SN Ia) 2007gi in the nearby galaxy NGC 4036. SN 2007gi is characterized by extremely high-velocity (HV) features of the intermediate-mass elements (Si, Ca, and S), with expansion velocities ($v_{\\rm exp}$) approaching $\\sim$15,500 km s$^{-1}$ near maximum brightness (compared to $\\sim$10,600 km s$^{-1}$ for SNe Ia with normal $v_{\\rm exp}$). SN 2007gi reached a $B$-band peak magnitude of 13.25$\\pm$0.04 mag with a decline rate of $\\Delta m_{15}(B)$(true) = 1.33$\\pm$0.09 mag. The $B$-band light curve of SN 2007gi demonstrated an interesting two-stage evolution during the nebular phase, with a decay rate of 1.16$\\pm$0.05 mag (100 days)$^{-1}$ during $t = 60$--90 days and 1.61$\\pm0.04$ mag (100 days)$^{-1}$ thereafter. Such a behavior was also observed in the HV SN Ia 2006X, and might be caused by the interaction between supernova ejecta and circumstellar material (CSM) around HV SNe Ia. Based on a sample of a dozen well-observed $R$-band (or unfiltered) light curves of SNe Ia, we confirm that the HV events may have a faster rise time to maximum than the ones with normal $v_{\\rm exp}$. ", "introduction": "Type Ia supernovae (SNe Ia) have been successfully utilized over the past decade to measure the cosmic expansion history \\citep{riess98, per99} and explore the nature of dark energy \\citep[e.g.,][]{riess04, riess07, ast06, wv07}. The foundation for the utility of SNe Ia as a cosmological tool is that some distance-independent observables, such as the light-curve shape parameters \\citep{phi93, riess95, jha07, per97}, the color parameters \\citep{wlf03, wxf05}, or both \\citep{tripp98, guy05}, have been found to correlate with their peak luminosity. These empirical correlations can be used to calibrate the luminosities of SNe Ia and measure their distances with a precision of $\\sim$9\\%. A recent result suggests that the luminosity standardization of SNe Ia can be improved to a level of $\\sim$6\\% by separating the SNe into two groups based on a spectroscopic criterion [for details, see \\citet{wxf09a}, hereafter W09]. The expansion velocity ($v_{\\rm exp}$) of the SN ejecta is inferred from the blueshift of the absorption minimum of Si`II $\\lambda$6355, and the SNe are divided into one group with normal $v_{\\rm exp}$ (hereafter ``Normal\" SNe Ia) and the other group with high $v_{\\rm exp}$ (hereafter ``HV\" SNe Ia)\\footnote{See \\citet{ben05} and \\citet{bran06} for a similar classification based on the velocity gradient and the strength of the absorption features, respectively.}. W09 found that the two groups have either different extinction laws or color evolution. The cause for such a dichotomy might be related to the properties of their progenitors. The HV SNe Ia are characterized by stronger absorption features of intermediate-mass elements (IMEs, such as Si, S, and Ca) at higher velocities as well as a red $B - V$ color around maximum brightness. SN 2002bo and SN 2006X are two of the best-studied examples of this class \\citep{ben04, wxf08a}. In particular, the interstellar Na I~D lines were found to show significant variations in the spectra of SN 2006X, likely pointing to the presence of CSM produced by the progenitor system \\citep{pat07, chu08}. The possible detection of CSM around SN 2006X is also supported by a flat evolution of the late-time light curve \\citep{wxf08a} and a detection of light echoes \\citep{wxf08b, cy08}. Similar variability of the Na I~D lines was also observed in SNe 1999cl and 2007le \\citep{blon09, sim09}, two other members of the HV SN Ia class. Contrasting with this, multi-epoch, high-resolution spectral observations of SN 2007af, a Normal SN Ia, do not reveal any significant signature of CSM absorption \\citep{sim07}. This raises the possibility that CSM might be preferentially present for the SNe Ia in the HV class. In this paper, we present optical observations of another member of the HV SN Ia class, SN 2007gi. Our goal is to increase the sample and understand the properties of well-observed HV SNe Ia. Observation and data reductions are described in \\S 2, while \\S 3 presents the $BVRI$ light curves, color curves, reddening estimate, and an analysis of the rise time. Section~4 presents the spectral evolution. Our discussions and conclusions are given in \\S 5. ", "conclusions": "In this paper, we present optical photometry and spectroscopy of SN 2007gi, a SN~Ia with an extremely high expansion velocity measured from IMEs (Si, Ca, and S). We also conduct a comparison study for a sample of SNe~Ia with high and normal expansion velocities (HV and Normal, respectively) to investigate differences in their photometric and spectroscopic behaviors. The $B$-band light curve of SN 2007gi shows a two-stage evolution: a decay rate of $\\beta = 1.16 \\pm 0.05$ mag (100~days)$^{-1}$ during $t \\approx 60$--90 days and $\\beta = 1.61 \\pm 0.05$ (100~days)$^{-1}$ thereafter. The latter decay rate is similar to those observed in Normal SNe Ia. This two-stage evolution is also present in the HV SN Ia 2006X. More late-time observations of the HV objects are necessary to establish whether this is a universal property. The $B - V$ color of SN 2007gi is found to evolve at a faster slope than that of Normal SNe Ia during the nebular phase, a trend that is also observed for the HV SN Ia 2006X \\citep{wxf08a} and most of the other HV events (Wang, X., et al. 2009c, in preparation). SN 2007gi was detected at $t \\approx 600$ days after $B$ maximum in {\\it HST}/WFPC2 images. The magnitudes and colors at this very late-time epoch suggest that SN 2007gi may not be contaminated by a light echo on interstellar scales. Using a dozen well-observed SNe Ia, we confirm the previous claim that HV SNe Ia tend to have a faster rise time than Normal SNe Ia \\citep{pig08}: at the same value of $\\Delta m_{15}$, the rise time of the HV objects is shorter than that of the Normal objects by 1--2 days. The spectral evolution of SN 2007gi is characterized by high expansion velocities measured from the absorption lines. The value of $v_{\\rm exp}$ is measured to be 15,500 km s$^{-1}$ at $t = -1$ day, $\\sim$50\\% higher than what is observed for Normal SNe Ia. A small flux ratio $\\mathcal{R}$(Si~II) = 0.07$\\pm$0.03 is observed, which traditionally means a hot photospheric temperature in the SN ejecta. Our synthetic spectral analysis using the SYNOW code, on the other hand, suggests that SN 2007gi has a typical photospheric temperature compared with other Normal SNe Ia. The two-stage evolution in the $B$-band light curve of SN 2007gi may suggest that an additional energy source besides radioactive decay plays a role in the nebular phase. Dust scattering (a light echo) seems unlikely because the tail luminosity would have remained nearly constant for a long time, which is not the case for SN 2007gi. Nevertheless, the possibility of a local light echo on a small scale (due to CSM dust) cannot be completely ruled out. Another potential energy source is interaction with CSM; detection of variable sodium lines in the spectra of some HV events provides some evidence for this scenario. Assuming that the CSM lies at a distance of $\\sim10^{17}$ cm from the supernova, it is expected that the outermost ejecta will begin to interact with the CSM at $\\sim$30 days after the $B$ maximum. This naturally accounts for a flatter evolution seen in the $B$-band light curve of the HV SNe Ia after $t = 40$ days. The difficulty with this scenario is the lack of convincing observational evidence for stripped material, such as low-velocity H$\\alpha$ emission in the nebular spectra [e.g., \\citet{mat05}, but see \\citet{leo08} for alternative explanations regarding the observed lack of hydrogen], and the lack of spectral evidence in support of ongoing CSM interaction. A possible explanation for the HV features observed in the HV SNe Ia is that there are density enhancements of the IMEs in the outer layers of the ejecta of the HV objects (due to a metallicity effect, delayed detonations, or CSM interaction); the $\\gamma$-ray heating is less effective, and the photosphere is formed at an outer layer with a higher expansion velocity compared to the Normal objects." }, "0910/0910.0915_arXiv.txt": { "abstract": "We present a pre-discovery $H$-band image of the \\object{HR 8799} planetary system that reveals all three planets in August 2007. The data were obtained with the Keck adaptive optics system, using angular differential imaging and a coronagraph. We confirm the physical association of all three planets, including HR~8799d, which had only been detected in 2008 images taken two months apart, and whose association with HR~8799 was least secure until now. We confirm that the planets are 2--3~mag fainter than field brown dwarfs of comparable near-infrared colors. We note that similar under-luminosity is characteristic of young substellar objects at the L/T spectral type transition, and is likely due to enhanced dust content and non-equilibrium CO/CH$_4$ chemistry in their atmospheres. Finally, we place an upper limit of $\\gtrsim$18~mag per square arc second on the $>$120~AU $H$-band dust-scattered light from the HR~8799 debris disk. The upper limit on the integrated scattered light flux is $10^{-4}$ times the photospheric level, 24 times fainter than the debris ring around HR~4796A. ", "introduction": "The three-planet system around the A5V star HR~8799 \\citep[][henceforth, M08]{marois_etal08b} is the first directly imaged extrasolar multi-planet system. Along with a reported extrasolar planet around Fomalhaut \\citep[A4V;][]{kalas_etal08}, HR~8799b, c, and d are also the first bona-fide extrasolar planets directly imaged around stars other than the Sun. The HR~8799 and Fomalhaut planetary systems share two important similarities. First, in both cases the planets are widely separated from their host stars, at projected separations between 24--119~AU. And second, both hosts are A stars surrounded by cold debris disks \\citep{aumann85, sadakane_nishida86}, circumscribing the planetary systems. The debris disk around HR~8799 is one of the most massive detected by {\\it IRAS} and is a factor of several brighter \\citep[$L_{\\rm IR}/L_\\ast=2.3\\times10^{-4}$;][]{moor_etal06, rhee_etal07} than that around Fomalhaut \\citep[$L_{\\rm IR}/L_\\ast=8\\times10^{-5}$;][and references therein]{rhee_etal07}. While Fomalhaut's disk has been spatially resolved at wavelengths ranging from the visible \\citep{kalas_etal05} to the submillimeter \\citep{holland_etal98, marsh_etal05}, the HR~8799 debris disk remained unresolved until recently \\citep[][henceforth, S09]{su_etal09}. Prior to the discovery of the planets around HR~8799 by M08, we targeted the star with the Keck adaptive optics (AO) system \\citep{wizinowich_etal00} to detect the debris disk in scattered light. The strength of its IR excess and a detection in the submillimeter \\citep[$\\approx$10~mJy at 850~$\\micron$;][]{williams_andrews06} indicated a substantial optical depth, which given suitable disk viewing geometry could result in a detection of dust-scattered light. The data presented here did not produce the sought-after scattered light detection, but have revealed all three known planets around HR~8799. In particular, we report the earliest-epoch image of the closest-in planet, HR~8799d. Prior epoch detections of HR~8799b and/or HR~8799c have been published in M08, \\citet{fukagawa_etal09}, and \\citet{lafreniere_etal09}. ", "conclusions": "We have presented an year-2007 image of all three known HR~8799 planets, including a detection of the inner-most planet HR~8799d that precedes its discovery in \\citet{marois_etal08b} by one year. Our data exclude the presence of additional $>$3~\\Mjup\\ outer planets between 68 and 160~AU from the star. We do not detect scattered light from the debris disk, and place an upper limit of $\\sim$18~mag~arcsec$^{-2}$ on its $H$-band surface brightness at 3--6$\\arcsec$ (120--240~AU) from the star. Based on a more robust $H$-band photometric calibration than was possible in \\citet{marois_etal08b}, we % confirm that HR~8799b, c, and d are very under-luminous or redder compared to field brown dwarfs of similar colors or intrinsic luminosities. We note that this is a distinctive feature of all young substellar objects near the L/T transition, and speculate that it may be due to enhanced dust content and a significant departure from CO/CH$_4$ chemical equilibrium in cool low-surface gravity substellar atmospheres." }, "0910/0910.0853_arXiv.txt": { "abstract": "We present the discovery of spectacular double X-ray tails associated with \\ga\\ and a possibly heated X-ray tail associated with \\gab, both late-type galaxies in the closest rich cluster Abell~3627. A deep \\chandra\\ observation of \\ga\\ allows us for the first time to examine the spatial and spectral properties of such X-ray tails in detail. Besides the known bright tail that extends to $\\sim$ 80 kpc from \\ga, a fainter and narrower secondary tail with a similar length was surprisingly revealed, as well as some intriguing substructures in the main tail. There is little temperature variation along both tails. The widths of the secondary tail and the greater part of the main tail also remain nearly constant with the distance from the galaxy. All these results challenge the current simulations. The \\chandra\\ data also reveal 19 X-ray point sources around the X-ray tails. We identified six X-ray point sources as candidates of intracluster ULXs with $L_{\\rm 0.3 - 10 keV}$ of up to 2.5$\\times10^{40}$ erg s$^{-1}$. \\gem\\ spectra of intracluster HII regions downstream of \\ga\\ are also presented, as well as the velocity map of these HII regions that shows the imprint of \\ga's disk rotation. For the first time, we unambiguously know that active star formation can happen in the cold ISM stripped by ICM ram pressure and it may contribute a significant amount of the intracluster light. We also report the discovery of a 40 kpc X-ray tail of another late-type galaxy in A3627, \\gab. Its X-ray tail seems hot, $\\sim$ 2 keV (compared to $\\sim$ 0.8 keV for \\ga's tails). The H$\\alpha$ data for \\gab\\ are also presented. We conclude that the high pressure environment around these two galaxies is important for their bright X-ray tails and the intracluster star formation. The soft X-ray tails can reveal a great deal of the thermal history of the stripped cold ISM in mixing with the hot ICM, which is discussed along with intracluster star formation. ", "introduction": "The intracluster medium (ICM) has long been proposed to play a vital role in galaxy evolution in clusters, through ram pressure and turbulent/viscous stripping of the galactic cold gas (e.g., Gunn \\& Gott 1972; Nulsen 1982; Quilis et al. 2000). As the halo gas and the cold interstellar medium (ISM) is depleted in the stripping process, the galactic star formation will eventually be shut down and blue disk galaxies may turn into red galaxies (e.g., Quilis et al. 2000). The removal of the cold ISM also affects the accretion history of the central SMBH. Stripping of the ISM of the disk galaxies in clusters has been extensively studied in simulations recently, with better resolution and more physics included (e.g., Abadi et al. 1999; Stevens et al. 1999; Quilis et al. 2000; Schulz \\& Struck 2001; Bekki \\& Couch 2003; Roediger \\& Hensler 2005; Kapferer et al. 2009). These simulations show that stripping has a significant impact on galaxy evolution (e.g., disk truncation, formation of flocculent arms, build-up of a central bulge and enhanced star formation in the inner disk at the early stage of the interaction). Besides the impact on galaxy evolution, another significant question related to stripping is the evolution of the stripped ISM. After the cold ISM is removed from the galaxy, the general wisdom is that the stripped gas is mixed with the ICM through evaporation eventually. However, it is now known that a fraction of the stripped ISM turns into new stars in the galactic halo or the intracluster space, as revealed from observations (e.g., Sun et al. 2007b, S07 hereafter) and simulations (e.g., Vollmer et al. 2001b; Schulz \\& Struck 2001; Kronberger et al. 2008; Kapferer et al. 2009). The stripped cold ISM, if it can survive long enough to reach the cluster center, can effectively heat the cluster core via ram-pressure drag, which makes gravitational heating by accretion a possible way to offset cooling (Dekel \\& Birnboim 2008 argued for 10$^{5}$ - 10$^{8}$ M$_{\\odot}$ clumps). Current HI observations of cluster spiral galaxies still fail to detect most of the HI gas missing from HI-deficient spiral galaxies in the intracluster space (e.g., Vollmer \\& Huchtmeier 2007), which implies that the bulk of the stripped HI gas has been heated out of the cold phase. However, little is known about the details of mixing, as the actual strength of heat conductivity and viscosity is poorly known. How quickly is the stripped cold ISM heated and mixed with the ICM? What observational signature will evaporation and mixing produce? What fraction of the stripped cold ISM turns into new stars? The mixing of the stripped cold ISM with the hot ICM will produce multi-phase gas. Depending on the poorly known details of mixing, prominent soft X-ray emission may be produced, as well as H$\\alpha$ emission. In this ``unified model'' for tails of cluster late-type galaxies, X-ray and H$\\alpha$ tails are simply manifestations of the cold ISM tail in the mixing process. Therefore, one can better understand mixing and the relevant micro-physics from multi-wavelength data, e.g., the amount of missing HI gas and the amount of the stripped gas in hotter phases. Obviously, a central problem is the energy transfer in the multi-phase gas, which is also a significant question for large cool cores in clusters. Moreover, the wake behind the galaxy produced by stripping provides a way to constrain the ICM viscosity (e.g., Sun et al. 2006, S06 hereafter; Roediger \\& Br$\\ddot{\\rm u}$ggen 2008a). Various dynamic features, e.g., bow shocks, tails and vortices, can be produced in stripping (e.g., Stevens et al. 1999; Schulz \\& Struck 2001; Roediger et al. 2006; Roediger \\& Br$\\ddot{\\rm u}$ggen 2008b, RB08 hereafter; Tonnesen \\& Bryan 2009). Observational evidence of stripping of cluster late-type galaxies is present in HI and H$\\alpha$ observations, either through HI deficiency or tails (e.g., Giovanelli \\& Haynes 1985; Gavazzi et al. 2001b; Kenney et al. 2004; Oosterloo \\& van Gorkom 2005; Chung et al. 2007; Yagi et al. 2007; S07). X-ray tails of late-type cluster galaxies have only been detected recently: C153 in A2125 at $z$=0.253 (Wang et al. 2004), and UGC~6697 in A1367 at $z$=0.022 (Sun \\& Vikhlinin 2005). However, there are only $\\sim$ 60 counts from C153 so the X-ray extension is not unambiguous (Wang et al. 2004). The \\xmm\\ data show that the X-ray tail of UGC~6697 may extend to $\\sim$ 90 kpc from the nucleus but most of the X-ray emission is within the galaxy. Moreover, the galaxy has a peculiar morphology and a tidal tail. Its complex kinematic behavior, revealed from the velocity map, implies the existence of a second galaxy hidden behind the main galaxy (Gavazzi et al. 2001a). Thus, UGC~6697 is a very complicated system where tidal interaction is important. There is also a wide tail (67 kpc$\\times$90 kpc) of the spiral NGC~6872 in the 0.5 keV Pavo group (Machacek et al. 2005), but it is uncertain how it was formed as the tail terminates on the dominant early-type galaxy of the group. The brightest known X-ray tail behind a cluster late-type galaxy is \\ga\\ in the closest rich cluster A3627 (S06), with $\\sim$ 80\\% of the X-ray emission beyond the galactic halo. One should be aware that strong X-ray tails of late-type galaxies are rare (e.g., none in Virgo, see Section 7.1). We have done a blind search of strong X-ray tails in nearby clusters ($z<0.06$) by extending our previous \\chandra\\ work (Sun et al. 2007a) to the current \\xmm\\ and \\chandra\\ archives. Despite extensive cluster data in the archives, e.g., mosaic fields of $\\sim$ 2 deg$^{2}$ around the Perseus and Coma clusters with \\xmm, only two more X-ray tails of cluster late-type galaxies were found. One is also in A3627 (\\gab) and will be discussed in this paper. The other one is NGC~4848 in the Coma cluster (Finoguenov et al. 2004) that was recently observed for 29 ks by \\chandra. However, both X-ray tails are shorter (40 - 50 kpc) and 2.4 - 5.5 times fainter than \\ga's tail. Thus, the proximity of \\ga, the high flux and the large length of its X-ray tail makes it the best target for detailed analysis and comparison with simulations. One should not confuse X-ray tails of early-type galaxies with those of late-type galaxies. Early-type galaxies in clusters have abundant X-ray ISM but little or no cold ISM. There are some X-ray tails of early-type galaxies reported (e.g., Machacek et al. 2006; Sun et al. 2007; Randall et al. 2009). Their X-ray tails are composed of the stripped hot ISM and do not co-exist with cold gas. A3627 is the closest massive cluster ($z$=0.0163, $\\sigma_{\\rm radial}$ = 925 km/s and $kT$=6 keV) rivaling Coma and Perseus in mass and galaxy content (Kraan-Korteweg et al. 1996; Woudt et al. 2008). It is also the sixth brightest cluster in RASS and the second brightest non-cool-core cluster after Coma (B$\\ddot{\\rm o}$hringer et al. 1996). \\ga\\ is a blue emission-line galaxy (Woudt et al. 2004) that is only $\\sim$ 180 kpc from the cluster's X-ray peak in projection. Its radial velocity (4680$\\pm$71 km~s$^{-1}$, Woudt et al. 2004) is close to the average velocity in A3627 (4871$\\pm$54 km~s$^{-1}$, Woudt et al. 2008) so most of its motion is probably in the plane of sky. If \\ga\\ is on a radial orbit, its real distance to the cluster center should be close to the projected distance. There is no sign of a galaxy merger and no strong tidal features around \\ga. S06 found a long X-ray tail behind \\ga, in both \\chandra\\ (14 ks) and \\xmm\\ data (MOS: 18 ks, PN: 12 ks). The tail extends to at least 70 kpc from the galaxy with a length-to-width ratio of $\\sim$ 10. The X-ray tail is luminous ($\\sim$ 10$^{41}$ erg s$^{-1}$), with an X-ray gas mass of $\\sim 10^{9}$ M$_{\\odot}$. S06 interpreted the tail as the stripped ISM of \\ga\\ mixed with the hot ICM. The \\chandra\\ data also reveal three hard X-ray point sources ($L_{X} \\sim 10^{40}$ erg s$^{-1}$) along the tail, and the possibility of all of them being background active galactic nuclei (AGNs) is very small. S06 suggested that some of them may be ultra-luminous X-ray sources (ULXs) born from active star formation in the tail. S07 further discovered a 40 kpc H$\\alpha$ tail and over 30 emission-line regions downstream of \\ga, up to 40 kpc from the galaxy. S07 concluded that they are giant HII regions in the halo downstream or in intracluster space. Sivanandam et al. (2010) observed \\ga\\ with IRAC and IRS on \\spi. A warm ($\\sim$ 160 K) molecular hydrogen tail was detected to at least 20 kpc from the galaxy from the IRS data, at the same position as the bright X-ray and H$\\alpha$ tail. The total mass of the warm H$_{2}$ gas is $\\sim 2.5\\times10^{7}$ M$_{\\odot}$. As the IRS fields only cover the front part of the X-ray tail, the above warm H$_{2}$ gas mass is only a lower limit. 8 $\\mu$m PAH emission is also detected from the bright HII regions identified by S07. The \\spi\\ results may imply a large reservoir of colder gas than the observed warm gas downstream of \\ga, coexisting with the hot ICM ($\\sim 7\\times10^{7}$ K) and the H$\\alpha$ emitting gas ($\\sim 10^{4}$ K). Obvious follow-ups include deep \\chandra\\ exposures and optical spectroscopic observations of the candidate HII regions. In this paper, we present the results from our deep \\chandra\\ observation of \\ga\\ and \\gem\\ GMOS spectroscopic observations. The plan of this paper is as follows: The \\chandra\\ and \\gem\\ observations and the data analyses are presented in Section 2. In Section 3, we discuss the X-ray tails of \\ga. Section 4 is on the properties of the surrounding ICM. Section 5 discusses the \\chandra\\ point sources and the intracluster HII regions. We also found another 40 kpc X-ray tail in A3627, associated with \\gab, which is discussed in Section 6. The results are discussed in Section 7 and Section 8 contains the summary. We adopt a cluster redshift of 0.0163 for A3627 (Woudt et al. 2008). Assuming H$_{0}$ = 71 km s$^{-1}$ Mpc$^{-1}$, $\\Omega$$_{\\rm M}$=0.27, and $\\Omega_{\\rm \\Lambda}$=0.73, the luminosity distance is 69.6 Mpc, and 1$''$=0.327 kpc. ", "conclusions": "\\subsection{X-ray Tails of Late-type Galaxies and Mixing} While the paucity of luminous X-ray tails of late-type galaxies in nearby clusters has been known (e.g., Sun et al. 2007a), the discovery of \\ga's double X-ray tails is a surprise. M86 also has double X-ray tails and Randall et al. (2008) interpreted it as the consequence of the aspherical potential of M86 and M86's orbit. M86 is an early-type galaxy so its potential structure is different from \\ga's and its tail is only composed of the stripped hot ISM. M86's tails are in fact connected to the ``plume'' on the north of the galaxy and are located at larger distance from the galaxy, instead of coming from the galaxy directly like \\ga's. Moreover, a spectacular H$\\alpha$ complex connecting M86 and the nearby disturbed spiral NGC~4438 was recently discovered by Kenney et al. (2008). Some H$\\alpha$ emission also extends to the position of M86's tail. Thus, tidal interaction may also be important for M86's X-ray tail. There also appears to be a split feature that resembles a double tail in NGC~4438's \\chandra\\ image (Machacek et al. 2004; Randall et al. 2008). However, one part of the split feature is almost along the distorted disk plane and tidal interaction should be important in this case (Machacek et al. 2004; Kenney et al. 2008). NGC~4438 is $\\sim$ 3 times more luminous than \\ga\\ in the $K_{\\rm s}$ band. Its truncated halo should be larger than \\ga's ($\\sim$ 15 kpc in radius, S07), also because the Virgo cluster is less massive than A3627. The features in NGC~4438 only extend to $\\sim$ 11 kpc from the nucleus, clearly within its halo. The other late-type galaxy in the proximity, NGC~4388, has many H$\\alpha$ filaments and a spectacular HI tail (Yoshida et al. 2002; Oosterloo \\& van Gorkom 2005). A faint X-ray extension can be traced to $\\sim$ 14 kpc from the nucleus in the direction of the HI tail, from the \\chandra\\ and \\xmm\\ data. There is also an X-ray extension to east in the disk plane, which creates the appearance of double tails (Randall et al. 2008). NGC~4388 is $\\sim$ 1.5 times more luminous than \\ga\\ in the $K_{\\rm s}$ band so similar to NGC~4438, its X-ray tail is still within its halo. Thus, we do not consider the one-sided X-ray features in NGC~4438 and NGC~4388 to be X-ray double tails based upon the current data. They are clearly much fainter than \\ga's tails, especially given \\ga's much larger distance ($\\sim$ 4.3 times) and the much higher absorption to the direction of \\ga. We can also define strong X-ray tails as one-sided features outside of the tidal truncation radius of the galactic halo (like the tails of \\ga\\ and \\gab), in order to distinguish them from the weak X-ray tails of NGC~4438 and NGC~4388. The observed X-ray emission of the tails is likely from the interfaces of the hot ICM and the cold stripped ISM, reflecting the multi-phase nature of mixing. Although the detailed mixing process that determines the emergent X-ray spectra is poorly understood, the observed tail temperature may be related to the mass-weighted temperature that is usually adopted in simulations, $T_{\\rm mw} = (M_{\\rm ISM} T_{\\rm ISM} + M_{\\rm ICM} T_{\\rm ICM}) / (M_{\\rm ISM} + M_{\\rm ICM}) \\approx T_{\\rm ICM} / (1+X)$, where $X = M_{\\rm ISM} / M_{\\rm ICM}$ and we assume that $T_{\\rm ISM} \\ll T_{\\rm ICM}$. For the X-ray tails of \\ga\\ and \\gab, $X = 2 - 6.5$ if we assume $T_{\\rm mw}$ is the observed spectroscopic temperature. However, it is easy to find out that the mass-weighted temperature is always larger than the spectroscopic temperature derived by fitting the observed spectra with a single-$kT$ model ($T_{\\rm spe}$). We examined this with the CEMEKL and APEC models. For the CEMEKL model, $d EM(T) \\propto (T/T_{\\rm max})^{\\alpha-1} d T$. Assuming an isobaric condition and for $\\alpha >$ -1, \\begin{equation} T_{\\rm mw} = \\frac{\\int T dm(T)}{\\int dm(T)} = \\frac{1+\\alpha}{2+\\alpha} T_{\\rm max} \\end{equation} For $\\alpha >$ -0.5, 1/3 $T_{\\rm max} < T_{\\rm mw} < T_{\\rm max}$. It is also reasonable to assume that $T_{\\rm max} \\leq T_{\\rm ICM}$. However, with a series of XSPEC simulations, we found that $T_{\\rm spe} < T_{\\rm mw}$. For example, $T_{\\rm spe}$ = 0.5 - 1 keV when -0.5 $< \\alpha <$ 0.5, while $T_{\\rm mw}$ = 2 - 3.6 keV for $T_{\\rm max}$ = 6 keV. This is not surprising as $T_{\\rm spe}$ is mainly determined by the centroid of the iron-L hump at low temperatures, which biases $T_{\\rm spe}$ low. As long as there are significant emission components at $kT$ = 0.4 - 2 keV, the iron-L hump will be strong to make $T_{\\rm spe}$ deviated from $T_{\\rm mw}$. On the other hand, the parameter $\\alpha$ in the CEMEKL model may reflect the age of the X-ray tail. Our simulations indeed show that $T_{\\rm spe}$ increases monotonously with $\\alpha$. From the CEMEKL fits to the spectra of \\ga\\ and \\gab, \\gab's stripping event may indeed be older. The above simple estimate implies that $T_{\\rm spe}$ is a certain fraction of the ICM temperature. The luminosity of the X-ray tail should also increase with the magnitude of the ambient ICM pressure, under an isobaric condition. Thus, the X-ray tails of late-type galaxies in groups and the outskirts of poor clusters may have temperatures (e.g., $<$ 0.4 keV) and densities too low to be bright in the \\chandra\\ and \\xmm\\ energy bands. This helps to explain the paucity of strong X-ray tails of late-type galaxies in nearby clusters (e.g., Sun et al. 2007a). Other factors include the generally high cluster background and the small number of late-type galaxies in the \\chandra\\ field of nearby clusters as the covered cluster volume is small. With the above scenario, the X-ray tail of a late-type galaxy is simply the manifestation of its cold ISM tail embedded in the hot ICM. Optical line emission is also expected and is indeed detected, e.g., the H$\\alpha$ tails of \\ga\\ and \\gab\\ (S07 and Section 6). HI observations of A3627 were made with Australia Telescope Compact Array (ATCA) in 1996 and no HI emission was detected from \\ga\\ or \\gab\\ (Vollmer et al. 2001a). The 3$\\sigma$ detection limit is $\\sim$ 3 mJy/beam in one velocity channel with a beamsize of 30$''$ (Vollmer et al. 2001a). Adjusting to the cluster distance used in this work and assuming a linewidth of 150 km/s (Vollmer et al. 2001a), the corresponding HI gas mass limit is $\\sim 5\\times10^{8}$ M$_{\\odot}$ per beam. While this limit is not weak for the remaining HI gas in the disks of \\ga\\ and \\gab\\ (roughly within one beam), the X-ray tails of \\ga\\ and \\gab\\ span over 7 - 15 times the beam size so a lot of diffuse HI gas can remain undetected downstream of the galaxy. In fact, with the sensitivity of the available ATCA data (an HI column density limit of 2$\\times10^{20}$ cm$^{-2}$), four of the seven HI tails in the Virgo cluster found by Chung et al. (2007) and the long HI tail of NGC~4388 (Oosterloo \\& van Gorkom 2005) would not be detected. Clearly deeper HI observations are required, although the nearby bright radio source PKS~1610-60 (43 Jy at the 1.4 GHz) makes the task difficult (8$'$ - 14$'$ from \\ga\\ and \\gab). Both \\ga\\ and \\gab\\ are expected to host a significant amount of atomic and molecular gas initially. From Blanton \\& Moustakas (2009), the total neutral plus molecular gas fraction (relative to the stellar mass) is $\\sim$ 70\\% for \\ga\\ and $\\sim$ 30\\% for \\gab. Assuming a type of Sc and a blue-band diameter of 20 kpc, the expected HI mass of \\ga\\ is $\\sim 2.4\\times10^{9}$ M$_{\\odot}$, from the empirical relation derived by Gavazzi et al. (2005). Similarly for \\gab\\ (a type of Sa and a blue-band diameter of 30 kpc), the expected HI mass is $\\sim 3.9\\times10^{9}$ M$_{\\odot}$. Thus, \\ga\\ is expected to have $\\sim (3-5)\\times10^{9}$ M$_{\\odot}$ of cold ISM initially, while \\gab\\ is expected to have $\\sim (4-9)\\times10^{9}$ M$_{\\odot}$ initially. These numbers can be compared with the X-ray gas mass in their X-ray tails (Table 2). Sivanandam et al. (2010) detected the mid-IR emission from warm H$_{2}$ gas ($\\sim$ 150 K) in the galaxy and the first 20 kpc of the main tail, with a total mass of $\\sim 2.5\\times10^{7}$ M$_{\\odot}$. The H$_{2}$ lines that IRS detects are rotationally excited lines that are generated very efficiently. The warm gas they probe is likely a small fraction of the total H$_{2}$ gas, if the bulk of the molecular gas is still cold. Deep HI and CO observations are required to recover the bulk of the cold gas in \\ga, \\gab\\ and their wakes. The same question can also be asked to other cluster late-type galaxies with signs of stripping. Does a stripped ISM tail show up in HI, H$\\alpha$ and X-rays simultaneously? We attempt to summarize the known examples of stripped tails for cluster late-type galaxies. A Venn diagram is plotted in Figure 15. While a lot of stripped tails are known for cluster late-type galaxies, there is currently little known overlap between different bands, especially between X-ray tails and HI tails. This may be due to the lack of sufficient data, either in X-rays (e.g., for the Virgo galaxies with HI tails, Chung et al. 2007) or in HI (e.g., the two galaxies in this paper). We are pursuing \\xmm\\ data for the Chung et al. (2007) galaxies and deeper ATCA data for the A3627 galaxies to better address this question. On the other hand, there are various reasons that A one-to-one correlation may not exist in every case, even with better data. Depending on its age and the surrounding ICM, an ISM tail may emit predominantly in the HI band or the X-ray band. Soft X-ray emission of the tail may only be bright in high-pressure environment. We also understand little about the details of mixing. It may be related to the mean free path of particles and the coherence length of the magnetic field, which would introduce a radial dependence on the efficiency of mixing. Nevertheless, we first need better data to update the Venn diagram shown in Figure 15. Deep X-ray data alone, like \\ga's, can reveal some thermal history of the stripped ISM. One surprise from \\ga's X-ray data is the constancy of the spectroscopic temperatures along both tails (Figure 7). Indeed extra care is required to interpret this spectroscopic temperature as it is mainly determined by the centroid of the combined iron-L humps from multi-$T$ gas. Nevertheless, the possible spectral difference between \\ga's tails and \\gab's tail is intriguing. This difference, if confirmed, combined with the temperature constancy of \\ga's tails, puts a constraint on the heating timescale of the stripped cold ISM. Clearly the knowledge of the HI/CO gas distribution in these two galaxies will help to tighten the constraint, although it is certain that this timescale depends on the ICM temperature. \\subsection{Comparison with Simulations} After the initial analytic work by Gunn \\& Gott (1972), many simulations have been run on the ram-pressure stripping of late-type galaxies in clusters (e.g., Abadi et al. 1999; Schulz \\& Struck 2001; Vollmer et al. 2001b; RB08; Kapferer et al. 2009; Tonnesen \\& Bryan 2009 and references in those papers). Recent efforts surpass the early ones in many respects, e.g., resolution, time steps, realistic treatment of the galaxy orbit and some important baryon physics (e.g., cooling and star formation). We mainly compare our results with recent simulations of three groups, the Innsbruck group with the GADGET-2/SPH code (Kronberger et al. 2008; Kapferer et al. 2009), the Bremen group with the FLASH/AMR code (Roediger et al. 2006; RB08) and the Columbia group with the {\\em Enzo}/AMR code (Tonnesen \\& Bryan 2009). The GADGET-2 simulations include recipes for cooling, star formation and stellar feedback. The model galaxies run through a wind tunnel with a constant pressure. The FLASH simulations are non-radiative and model the flight of a disk galaxy through a galaxy cluster with realistic ICM and potential distributions. The {\\em Enzo} simulations include cooling but not feedback (Tonnesen \\& Bryan 2009). To mimic effects of some heating processes not in simulations, they impose a minimum temperature floor. The model galaxy is also subject to a constant ram pressure. The inclusion of cooling is important, as it indeed happens as revealed from our data (the intracluster HII regions and star clusters) and it has profound effects in simulations (Kapferer et al. 2009; Tonnesen \\& Bryan 2009). Cooling also naturally adjusts the ISM distribution on the disk plane, creating multi-phase ISM with enhancement and holes (Kapferer et al. 2009; Tonnesen \\& Bryan 2009), which is not present in the non-radiative runs. However, cooling needs to be offsetted by stellar feedback. Unknown efficiencies of transport processes (viscosity and heat conduction) because of unclear strength and configuration of magnetic field, plus the other uncertainties in the prescriptions of star formation and stellar feedback, make the tasks difficult (e.g., Tonnesen \\& Bryan 2009). We emphasize that comparison at the current stage is not straightforward as galaxies in these simulations are always more massive than \\ga\\ and high pressure environment was not explored in all simulations. We also caution that it is best to compare the observational data with the mock data generated from simulations. Kapferer et al. (2009) and Tonnesen \\& Bryan (2009) attempted to create X-ray mock data but more work is required, e.g., folding with the actual \\chandra\\ response, projection on the ICM background and the use of spectroscopic temperature rather than mass-weighted temperature (Section 7.1). Tails generally appear longer and wider in simulations than in observations. The simulations by Kapferer et al. (2009) show that the X-ray tails can be traced to $\\sim$ 300 kpc from a galaxy that is about 4 times more massive than \\ga, with an ambient pressure similar to \\ga's. The simulated X-ray tails are also very clumpy, with widths generally increasing with the distance from the galaxy. The FLASH simulations also produce longer tails than observed in this work, but are difficult to compare with the X-ray data directly. As RB08 pointed out, the observed tail length depends on the mass-loss per orbital length, which is the highest in poor clusters as galaxies in rich clusters move too fast and the ram pressure in groups is not high enough. The FLASH simulations also show tail flaring (i.e. increasing width with distance from the galaxy downstream) in all cases. The velocity perpendicular to the direction of the ram pressure comes from the random motion of the ICM, the disk rotation and turbulence in the wake. For a rotation velocity of 100 km/s over 50 Myr, the offset is 5 kpc, which is not small in \\ga's case as its tails are narrow. The Tonnesen \\& Bryan (2009) simulations produce long but narrower tails. The tails are also very clumpy. None of these simulations can reproduce the double tails of \\ga. The simulated tails are generally too clumpy with strong turbulence. This simple comparison may point to a higher ICM viscosity than present in the simulations. The Reynold number is $\\sim 3{\\cal M}(L/\\lambda)$, where ${\\cal M}$ is the Mach number, $L$ is the size of the remaining galactic gas disk and $\\lambda$ is the mean free path of particles in the ICM. For unmagnetized gas, $\\lambda \\sim$ 10 kpc around \\ga. The radius of the remaining H$\\alpha$ disk and the galactic X-ray emission is only $\\sim$ 1.7 kpc (S07). Thus, the Reynold number is on the order of unity which implies a laminar flow. Of course the presence of magnetic field would increase the Reynold number. In the simulations by Kapferer et al. (2009), a bright spot in the X-ray images is always found at distances of a few tens of kpc from the galaxy, caused by compressional heating. This feature can be compared with the bright spot in the main tail (Figure 1 and 5). However, the location of this feature is generally too far from the galaxy in simulations with high ambient pressure. Such a bright spot is also not observed in the secondary tail. The most obvious question about \\ga's X-ray tails is why are there double tails. The simplest explanation is to involve another dwarf galaxy forming a pair with \\ga. This was in fact simulated by Kapferer et al. (2008), showing double gaseous tails, stellar tidal tails and a stellar bridge between two galaxies. The X-ray gas mass of the secondary tail is 40\\% - 50\\% of the mass of the main tail, assuming the same filling factors of the X-ray gas. If we simply take the ratio of their X-ray gas masses as the ratio of their ISM gas masses, the stellar mass of this assumed dwarf that is responsible for the secondary tail should be $\\sim$ 30\\% of \\ga's stellar mass, from the ISM mass - stellar mass relation summarized by Blanton \\& Moustakas (2009). Such a galaxy should be easily detected, at least through the tidal tails and bridge it causes even if it hides behind \\ga. However, the SOAR images do not show the presence of another galaxy near \\ga. There are also no tidal features upstream of the galaxy. There are some continuum features downstream of the galaxy as shown in S07. However, all of them are around confirmed HII regions and have blue colors. Our new \\hst\\ data only reinforce the conclusion from the SOAR images. The central region of the galaxy is very dusty but the galaxy appears regular without significant tidal features in the \\hst\\ $I$ band (Paper IV, in preparation). Thus, we conclude that both X-ray tails originate from \\ga. The discovery of two separated stripping tails from one galaxy presents a challenge to our understanding of stripping. In simulations (e.g., RB08; Kapferer et al. 2009; Tonnesen \\& Bryan 2009), cold ISM is removed from the sides of the disk or through holes and is quickly mixed to form one clumpy tail because of disk rotation, momentum transferred from the ICM and turbulence in the wake, especially if projected on the plane of sky. Perhaps we should begin to consider more realistic ISM distributions in the galaxy. Optical images clearly show the presence of two thick spiral arms, one to the north and the other one to the south (S07; better shown in the \\hst\\ images). Maybe the two X-ray tails correspond to stripping of molecular gas concentrated around these two spiral arms. CO images of nearby face-on spiral galaxies indeed show that molecular gas distribution follows major spiral arms in many cases (e.g., Kuno et al. 2007). This is also generally true for HI gas (e.g., Adler \\& Westpfahl 1996). For \\ga, we present a simple cartoon in Figure 16. The north arm is apparently the trailing arm where stripping is easier so a brighter X-ray tail is produced. The south arm is the leading arm so gas removal there is more difficult as most gas would be pushed inwards to deeper galactic potential first. This also explains the fact that more intracluster HII regions (at $>$ 15 kpc from the nucleus) are found around the main tail while the numbers of HII regions in the halo are similar at two sides. The curvature of the secondary tail may come from the galactic rotation, which is high enough to cause the observed offset. The main tail may also be curved in 3D but appears straight in projection. The rather regular wake structure may imply significant viscosity. On the other hand, draping of the ICM magnetic field (e.g., Dursi \\& Pfrommer 2008) can also inhibit the motion of the stripped ISM perpendicular to the direction of ram pressure. In summary, it remains to be seen whether simulations can reproduce the double X-ray tails of \\ga\\ as observed, and the gas properties in other bands (H$\\alpha$ and H$_{2}$ at least). We also emphasize that the comparison between data and simulations is not trivial and straightforward. Observationally we need deeper data at multiple wavelengths. For simulations, mock data need to be produced with sensitivities matching the actual observations. Uncertainties in the viscosity, heat conduction, cooling and stellar feedback should be explored with different runs. Eventually we need to compare the multi-wavelength data (HI, CO, IR, optical and X-rays) with the mock data in the same band. Besides the simple comparison of morphology, a lot of details of the tail properties (e.g., temperature distribution, filling factors of the gas in different phases and energy transfer between gas in different phases) can be better examined. \\subsection{Formation of Young Stars and X-ray Binaries in the Stripped ISM} The new observations of \\ga, for the first time, unambiguously confirm the active star formation in the cold ISM stripped by the ICM ram pressure. As emphasized in S07, \\ga's immediate surroundings are devoid of bright galaxies. There are only two galaxies that are within 4 mag of \\ga\\ in the $I$ band, within 100 kpc from the end of \\ga's X-ray tail (G1 and G2 in Figure 1 of S07). Their redshifts are unknown and they are 1.9 - 2.7 mag fainter than \\ga\\ in the $I$ band. There is also no such galaxy within 70 kpc from the nucleus of \\ga. Both G1 and G2 are undisturbed and there are no tidal features between \\ga\\ and G1 (or G2). Our \\hst\\ data do not reveal another dwarf galaxy near \\ga\\ (Paper IV, in preparation). The velocity map of the HII regions also shows no extra component, besides \\ga's rotation pattern. It is also known that the cluster potential is not the main factor for stripping (S07; Sivanandam et al. 2010), although it indeed affects the size of \\ga's dark matter halo. Thus, we conclude that ICM ram pressure is responsible for the stripping of the ISM. In our case, the ambient pressure is even high enough to remove some molecular gas from the galaxy. The HI gas has typical pressures of $P/k_{\\rm B} \\sim 10^{3} - 10^{4}$ K cm$^{-3}$, while the pressure in the core of a starless giant molecular cloud can be higher than $10^{6}$ K cm$^{-3}$. The ambient total thermal pressure (electron + ions) is 1.3$\\times10^{5}$ K cm$^{-3}$ for $n_{\\rm e}$ = 10$^{-3}$ cm$^{-3}$ and $kT$=6 keV, while the ram pressure is 3.2$\\times10^{5}$ K cm$^{-3}$ for $n_{\\rm e}$ = 10$^{-3}$ cm$^{-3}$ and $v_{\\rm gal}$ = 1500 km/s. After removing the cold ISM from the galaxy, the key for the intracluster star formation is that the stripped ISM must be able to cool, likely aided by shocks and the absence of the UV radiation field that is strong on the disk plane. Intracluster star formation in the stripped ISM was first suggested in simulations (Schulz \\& Struck 2001; Vollmer et al. 2001b). Schulz \\& Struck (2001) even suggested formation of dwarf galaxies in this mode. In the new GADGET-2/SPH simulations by Kronberger et al. (2008) and Kapferer et al. (2009), new stars are formed in the stripped ISM, forming $\\sim$ 1 kpc structures with masses of up to 10$^{7}$ M$_{\\odot}$ individually to hundreds of kpc downstream of the galaxy. However, all simulations suffer from uncertainties in transport processes like viscosity and heat conduction, which are important for determining the heating rate. More evidence of intracluster star formation, either HII regions or blue star clusters, has been revealed recently (e.g., Lee et al. 2000; Gerhard et al. 2002; Cortese et al. 2004; Cortese et al. 2007; Yoshida et al. 2008). They are only at one side of the galaxy. In some cases, tidal interaction may be important (e.g., Lee et al. 2000). But they all lack unambiguous associations with gaseous tails. We especially highlight the growing number of examples of one-sided blue stellar filaments and star clusters behind cluster galaxies, which includes two galaxies in A2667 and A1689 at $z \\sim$ 0.2 (Cortese et al. 2007) and a dwarf galaxy RB~199 in the Coma cluster (Yoshida et al. 2008). The connection between the first two galaxies at $z \\sim 0.2$ with \\ga\\ was discussed in S07. In the case of RB~199, Yoshida et al. (2008) found narrow blue filaments, knots, H$\\alpha$ filaments and clouds extending to 80 kpc from the galaxy in one side. It is a rather remarkable case as the galaxy is much fainter than all previous galaxies, with a stellar mass of $\\sim 8 \\times 10^{8}$ M$_{\\odot}$. The stellar mass of the downstream structures is $\\sim 10^{8}$ M$_{\\odot}$, which is $\\sim$ 10\\% of the galactic mass. Yoshida et al. (2008) also favored a mechanism similar to what we suggested for \\ga\\ in S07. These three examples may imply a rather common evolution stage for cluster late-type galaxies. As the stage with bright blue stellar trails is short ($\\leq$ 200 Myr), more similar examples are expected only in large surveys. If we assume that intracluster star formation happens in the stripped ISM of every cluster spiral, can this mode contribute a significant fraction of the intracluster light? The answer depends on the mass fraction of intracluster stars formed out of the stripped ISM (or the star formation efficiency). S07 gave a rough estimate of $\\sim$ 1\\%, which may only be a lower estimate as blue stellar filaments and clusters (not shining in H$\\alpha$) are not counted. While a detailed account of the total stellar mass downstream of \\ga\\ is the aim of Paper IV, here we take 1\\% as a lower limit. If the efficiency of the intracluster star formation is only 1\\%, the intracluster light contributed from star formation in stripped ISM is small, given the mass fraction of the intracluster light to the total stellar light in clusters ($\\sim$ 20\\%, likely depending on the halo mass, e.g., Krick \\& Bernstein 2007). However, the case of RB~199 implies a fraction of $\\sim$ 10\\%, if the original ISM in RB~199 had a mass comparable to its stellar mass (e.g., Blanton \\& Moustakas 2009). In the simulations by Kronberger et al. (2008) and Kapferer et al. (2009), 7\\% - 12\\% of the initial ISM turns into new stars in the wake. If an efficiency of 10\\% is adopted, the intracluster light contributed by ram pressure stripping is a significant fraction of the total intracluster light. However, this efficiency may not be too high given the star formation efficiency in galaxies. In galactic disks, $\\sim$ 1\\% of gas is converted to stars per free-fall time (e.g., Leroy et al. 2008). In 100 Myr (a typical orbital time in the disks), $\\sim$ 6\\% of gas is converted to stars (e.g., Kennicutt 1998; Leroy et al. 2008). We note that 100 Myr is also about the orbital time of the residual H$\\alpha$ disk ($\\sim$ 1.7 kpc radius with an orbital velocity of $\\sim$ 100 km/s) and close to the age of the X-ray tails. Clearly, a lot of work is required to constrain the intracluster star formation efficiency. The efficiency may also change with the mass of the host galaxy as more stars are formed inside the halo of more massive galaxies. In principle, the ratio of the intracluster SNIa to the other types can also inform us the significance of intracluster star formation. As active star formation is confirmed downstream of the galaxy, the discovery of ULXs is not a surprise. As we estimated in Section 5.2, the total luminosity of ULXs can be consistent with the expectation from the intracluster star formation rate. The space density of X-ray point sources also traces the strength of the star formation, stronger in the halo and weaker around the unbound tail. A similar system may be the high-velocity system of NGC~1275. Gonzalez-Martin et al. (2006) found a concentration of eight ULXs ($L_{\\rm 0.5-7 keV} > 3\\times10^{39}$ erg s$^{-1}$) around the high-velocity system. They may be associated with the active star formation in and downstream of the high-velocity system, similar to what is happening in \\ga." }, "0910/0910.0254_arXiv.txt": { "abstract": "Even though the existence of intermediate-mass black holes (IMBHs, black holes with masses ranging between $10^{2-4}\\,M_{\\odot}$) has not yet been corroborated observationally, these objects are of high interest for astrophysics. Our understanding of the formation and evolution of supermassive black holes (SMBHs), as well as galaxy evolution modeling and cosmography would dramatically change if an IMBH were to be observed. From a point of view of traditional photon-based astronomy, {which relies on the monitoring of innermost stellar kinematics}, the {\\em direct} detection of an IMBH seems to be rather far in the future. However, the prospect of the detection and characterization of an IMBH has good chances in lower-frequency gravitational-wave (GW) astrophysics using ground-based detectors such as LIGO, Virgo and the future Einstein Telescope (ET). We present an analysis of the signal of a system of a binary of IMBHs (BBH from now onwards) based on a waveform model obtained with numerical relativity simulations coupled with post-Newtonian calculations at the highest available order. IMBH binaries with total masses between $200-20000\\,M_\\odot$ would produce significant signal-to-noise ratios (SNRs) in advanced LIGO and Virgo and the ET. We have computed the expected event rate of IMBH binary coalescences for different configurations of the binary, finding interesting values that depend on the spin of the IMBHs. The prospects for IMBH detection and characterization with ground-based GW observatories would not only provide us with a robust test of general relativity, but would also corroborate the existence of these systems. Such detections should allow astrophysicists to probe the stellar environments of IMBHs and their formation processes. ", "introduction": "\\label{sec.motivation} By following the stellar dynamics at the center of our Galaxy, we have now the most well-established evidence for the existence of a SMBH. The close examination of the Keplerian orbits of the so-called S-stars (also called S0-stars, where the letter ``S'' stands simply for source) has revealed the nature of the central dark object located at the Galactic Center. By following S2 (S02), the mass of SgrA$^*$ was estimated to be about $3.7\\times 10^6\\,M_{\\odot}$ within a volume with radius no larger than 6.25 light-hours \\citep{SchoedelEtAl03,GhezEtAl03b}. More recent data based on 16 years of observations set the mass of the central SMBH to $\\sim 4 \\times 10^{6} \\, M_{\\odot}$ \\citep{EisenhauerEtAl05,GhezEtAl05,GhezEtAl08,GillessenEtAl09}. Massive black holes in a lower range of masses may exist in smaller stellar systems such as globular clusters. These are called intermediate-mass black holes (IMBHs) because their masses range between $M\\sim 10^{2}$ and $M\\sim 10^{4}\\,M_\\odot$, assuming that they follow the observed correlations between SMBHs and their host stellar environments. Nevertheless, the existence of IMBHs has never been confirmed, though we have some evidence that could favor their existence \\citep[see][ and references therein]{MillerColbert04,Miller09}. If we wanted to apply the same technique to detect IMBHs in globular clusters as we use for SMBHs in galactic centers, we would need ultra-precise astronomy, since the sphere of influence of an IMBH is {$\\sim$ few arc seconds. For instance, for a $10^4\\,M_{\\odot}$ IMBH, the influence radius is of $\\sim 5''$ assuming a central velocity dispersion of $\\sigma = 20\\,{\\rm km\\,s}^{-1}$ and a distance of $\\sim 5$ kpc.} The number of stars enclosed in that volume is only a few. Currently, with adaptive optics, one can aspire --optimistically-- to have a couple of measurements of velocities if the target is about $\\sim 5$ kpc away on a time scale of about 10 yrs. The measures depend on a number of factors, such as the required availability of a bright reference star in order to have a good astrometric reference system. Also, the sensitivity limits correspond to a K-band magnitude of $\\sim$ 15, (B- MS stars at 8 kpc, like e.g. S2 in our Galactic Center). This means that, in order to detect an IMBH or, at least, a massive dark object in a globular cluster center by following the stellar dynamics around it, one has to have recourse to the Very Large Telescope interferometer and to one of the next-generation instruments, the VSI or GRAVITY \\citep{GillessenEtAl06,EisenhauerEtAl08}. In this case we can hope to improve the astrometric accuracy by a factor of $\\sim 10$. Only in that scenario would we be in the position of following closely the kinematics around a potential IMBH so as to determine its mass. GW astronomy could contribute to IMBH detection. In the past years, the field has reached a milestone with the construction of an international network of GW interferometers that have achieved or are close to achieving their design sensitivity. Moreover, the first-generation ground-based detectors LIGO and Virgo will undergo major technical upgrades in the next five years that will increase the volume of the observable universe by a factor of a thousand \\footnote{\\scriptsize{http://www.ligo.caltech.edu/advLIGO/, http://wwwcascina.virgo.infn.it/advirgo/}}. The data that will be taken by the advanced interferometers are expected to transform the field from GW detection to GW astrophysics. The availability of accurate waveform models for the full BBH coalescence in order to construct templates for match-filtering is crucial in the GW searches for compact binaries. The construction of this kind of templates has recently been made possible thanks to the combination of post-Newtonian calculations of the BBH inspiral and numerical relativity simulations of the merger and ringdown. Two approaches to this problem are the effective-one-body techniques~\\citep{Buonanno:1998gg,Buonanno:2009qa} and the phenomenological matching of PN and NR results~\\citep{Ajith:2007kx,Ajith:2009bn,Santamaria:SpinTemp}. In this paper we use the results of the latter. The structure of this paper is as follows: In section~\\ref{sec:astro} we expand the astrophysical context to this problem and give a description of the different efforts made to address the evolution of a BBH in a stellar cluster from its birth to its final coalescence. In section~\\ref{sec:wfmodel} we introduce the techniques used in the data analysis of the waveform modeling of BBH coalescences, present our hybrid waveform model, and use the model to compute and discuss expected signal-to-noise ratios in present and future GW detectors. The use of the new waveform model allows us to give an improved estimate for the number of events one can expect for several physical configurations in Advanced LIGO and the Einstein Telescope, which we present in section~\\ref{sec:events}. We conclude with a summary of our results and future prospects of our work in section~\\ref{sec:concl}. ", "conclusions": "\\label{sec:concl} Even though we do not have any evidence of IMBHs so far, a number of theoretical works have addressed their formation in dense stellar clusters. If we were to follow the same techniques that have led us to discover the SMBH in our own Galaxy, we would need the Very Large Telescope interferometer and next-generation instruments, such as the VSI or GRAVITY, which should be operative in the next $\\sim 10$ yrs. An alternative or even complementary way of discovering IMBHs is via their emission of GWs when they are in a BBH system. The identification and characterization of these systems rely on accurate waveform modeling of their GW emission, which has been made possible by the success of numerical relativity in simulating the last orbits of the BBH coalescence and the coupling of these results to analytical post-Newtonian calculations of the inspiral phase. We use a PN-NR hybrid waveform model of the BBH coalescence based on a construction procedure in the frequency domain \\citep[see][for details]{Santamaria:SpinTemp}. Using this hybrid waveform, we have estimated the SNR corresponding to the current and Advanced LIGO and Virgo detectors, the proposed ET and the space-based LISA at a distance of $z=1$, i.e. 6.68 Gpc. The results indicate that IMBH binaries will produce SNRs sufficient for detection in advanced LIGO and Virgo and notably larger SNRs in the ET, thus making them interesting sources to follow up on. Eventual observations of IMBH binaries with future ground-based detectors could be complementary to those of LISA, which is expected to detect these systems with moderate SNRs and to be more sensitive to SMBH binaries instead. More remarkably, in principle, if LISA and the ET are operative at the same time, they could complement each other and be used to track a particular event. Furthermore, we have revisited the event rate of BBHs for various detectors and find encouraging results, within the inherent uncertainties of the approach. Our estimations are consistent with previous works, and additionally we have quantified the effect of the total spin of the binary in the expected event rates. We have estimated the distance to which Advanced LIGO and the ET will be able to see binaries of IMBHs. This quantity depends strongly on the mass ratio and spins of the binary. For Advanced LIGO, equal-mass, non-spinning configurations of observed total mass $\\sim 200$--$700\\,M_\\odot$ can be seen up to $z \\sim 0.8$. If the spins are aligned with the total angular momentum and significant ($\\chi_{1,2} \\sim 0.75$), the reach increases to $z \\sim 1.5$ for observed total masses of $\\sim 400\\,M_\\odot$. The ET will be able to explore even more remote distances, reaching to $z\\sim 5$ and further. Our present knowledge of stellar formation at such large redshifts is incomplete, therefore we have computed the event rates for the ET integrating only until $z=5$. We have compared three star formation models and three different configurations of the binary. The effect of the particular formation model is not very significant; remarkably, the physical configuration does however influence the final rates. We provide the results for three physical configurations, taking into account both single- and double-cluster channels in the binary formation. For a fully correct calculation of the event rate integrated over all possible configurations, more detailed knowledge of the distribution of spins and mass ratios of IMBH binaries formed in globular cluster is required. Advanced ground-based detectors are designed to be able to operate in different modes so that their sensitivity can be tuned to various kinds of astrophysical objects. Considering the importance of an eventual detection of a BBH, the design of an optimized Advanced LIGO configuration for systems with $M \\sim 10^{2-4}\\,M_{\\odot}$ would be desirable to increase the possibility of observing such a system. In case an IMBH binary coalescence were detected, the recovery and study of the physical parameters of the system could serve to test general relativity and prove or reject alternative theories, such as scalar-tensor type or massive graviton theories. The {\\em direct} identification of an IMBH with GWs will be a revolutionary event not only due to the uncertainty that surrounds their existence and their potential role in testing general relativity. The information encoded in the detection will provide us with a detailed description of the environment of the BBH/IMBH itself." }, "0910/0910.5186_arXiv.txt": { "abstract": "% We report a chromospheric jet lasting for more than 1~hr observed by {\\it Hinode} Solar Optical Telescope in unprecedented detail. The ejection occurred in three episodes separated by 12--14~min, with the amount and velocity of material decreasing with time. The upward velocities range from 438 to $33 \\kmps$, while the downward velocities of the material falling back have smaller values (mean: $-56 \\kmps$) and a narrower distribution (standard deviation: $14 \\kmps$). The average acceleration inferred from parabolic space-time tracks is $141 \\m \\, {\\rm s^{-2}}$, a fraction of the solar gravitational acceleration. The jet consists of fine threads ($0\\farcs5$--$2\\arcsec$ wide), which exhibit coherent, oscillatory transverse motions perpendicular to the jet axis and about a common equilibrium position. These motions propagate upward along the jet, with the maximum phase speed of $744 \\pm 11 \\kmps$ at the leading front of the jet. The transverse oscillation velocities range from $151$ to $26 \\kmps$, amplitudes from $6.0$ to $1.9 \\Mm$, and periods from $250$ to $536 \\s$. The oscillations slow down with time and cease when the material starts to fall back. The falling material travels along almost straight lines in the original direction of ascent, showing no transverse motions. These observations are consistent with the scenario that the jet involves untwisting helical threads, which rotate about the axis of a single large cylinder and shed magnetic helicity into the upper atmosphere. ", "introduction": "\\label{sect_intro} Transient, small-scale ejections of plasma from the lower atmosphere are common manifestations of solar activity. Cool plasma ejections were observed as emission in \\Ha or absorption at other wavelengths and historically called {\\it surges} \\citep{NewtonH.surge-discovery.1934MNRAS..94..472N}. Hot ejections were usually called {\\it jets} and observed as emission in ultraviolet \\citep{Brueckner.Bartoe.UVjet1983ApJ...272..329B}, extreme ultraviolet \\citep[EUV;][]{Alexander.Fletcher.jet.1999SoPh..190..167A}, and soft X-rays \\citep{ShibataK.1st-SXT-jet.1992PASJ...44L.173S, StrongK.SXT-jet1992PASJ...44L.161S}. Some observations \\citep{SchmiederB.cool-surge.hot-jet1988A&A...201..327S, ChaeQiu.jet.surge1999ApJ...513L..75C, JiangYC.surge.jet.twist2007A&A...469..331J} have indicated that cool surges and hot jets are closely related in space and time, but their physical relationship has not been established.\t% Torsional motions or helical features have long been seen in surges or jets \\citep[e.g.,][]{XuAA.surge-rotate.1984AcASn..25..119X, Kurokawa.untwist-filamt1987SoPh..108..251K, ShimojoM.jet-stat.1996PASJ...48..123S, Patsourakos.EUVI-jet2008ApJ...680L..73P}, while their exact causes remain unclear. From now on, we refer to both surges and jets with a general term {\\it jets}, unless otherwise noted. Solar jets are commonly associated with\t\t% flux emergence \\citep{Roy.surge.magnetic.1973SoPh...28...95R}\t% or moving magnetic features \\citep{Gaizauskas.movingBfield.surge1982AdSpR...2...11G, BrooksDH.surge2007ApJ...656.1197B}. Models involving magnetic reconnection have thus been proposed \t% \\citep[e.g.,][]{RustD.surge-reconn.1968IAUS...35...77R, YokoyamShibata.jetModel1995Natur.375...42Y}. Mechanisms \\citep[see review in][]{Canfield.surge-jet1996ApJ...464.1016C}\t\t% suggested for accelerating jet material to 10--1000~$\\kmps$ include reconnection outflows driven by the slingshot effect of magnetic tension \\citep{YokoyamShibata.jetModel1995Natur.375...42Y}, the pressure gradient behind the shock formed by reconnection outflows \\citep{YokoyamShibata.jetModel1996PASJ...48..353Y, Tarbell.Ryutova.jet.1999ApJ...514L..47T}, chromospheric evaporation caused by heating from the associated flare \\citep{ShimojoM.jet-phys.2000ApJ...542.1100S}, and relaxation of magnetic twists \\citep{Shibata.Uchida.helic-jetMHD.1985PASJ...37...31S}. \\hinode \\citep{Kosugi.Hinode2007SoPh..243....3K}, with its superior resolutions, has provided unprecedented details and spurred renewed interest in solar jets. They were found to be ubiquitous on various spatial and temporal scales in the \\ion{Ca}{2} H line \\citep{Shibata.CaHjet.2007Sci...318.1591S, Nishizuka.giantCaHjet.2008ApJ...683L..83N}, EUV \\citep{Culhane.EIS-jet2007PASJ...59S.751C, Moreno-Insertis.EISjet2008ApJ...673L.211M}, and soft X-rays \\citep{Savcheva.XRT.jet.stat2007PASJ...59S.771S, NittaN.jetHe3.2008ApJ...675L.125N}. In this Letter, we report an intriguing jet observed by \\hinode in great detail, and attempt to obtain new clues to some unanswered questions mentioned above. \\begin{figure*}[thb] % \\epsscale{1.2}\t% \\plotone{f1.eps} \\caption[]{ Negative \\hinode \\ion{Ca}{2} H images.\t\t% The diagonal dotted line \t% marks the jet axis. The dashed box indicates a $10\\arcsec$ wide cut along the axis for the time-distance diagram shown in Figure~\\ref{para.eps}(a),\t\t% while the numbered dotted lines represent $1\\arcsec$ narrow cuts perpendicular to the axis (see online movie~1). } \\label{maps.eps} \\end{figure*} ", "conclusions": "% \\label{sect_conclude} We have presented kinematic measurements of both axial and transverse motions of a chromospheric jet observed by \\hinode SOT at high spatial and temporal resolutions. This study complements previous low resolution counterparts of \\Ha surges \t% and EUV or X-ray jets, and offers new insights to this type of phenomena. Our major results and interpretations are as follows. \\begin{enumerate} \\item\t% The ejection occurs in three episodes separated by 12--14~min, rather than continuously. The amount and velocity of ejected material decrease with time.\t% The ejecting velocities have a wide range from 438 to $33 \\kmps$, while the velocities of material falling back \t\t% have a narrow range (mean: $-56 \\kmps$ and standard deviation: $14 \\kmps$). The acceleration inferred from parabolic tracks in the time-distance diagram has a mean of $141 \\m \\, {\\rm s^{-2}}$, a fraction of the solar gravitational acceleration. \\item\t% The jet consists of $0\\farcs5$--$2 \\arcsec$ thick fine threads, which exhibit oscillatory transverse motions across the jet about a common equilibrium position. These oscillations have velocities ranging from $151 \\pm 6$ to $26 \\pm 3 \\kmps$, amplitudes from $6.0 \\pm 0.2$ to $1.9 \\pm 0.2 \\Mm$, and periods from $250 \\pm 6$ to $536 \\pm 19 \\s$. The upward propagation of the oscillations has a maximum phase speed of $744 \\pm 11 \\kmps$ (comparable to the coronal Alfv\\'{e}n speed) associated with the leading front of the jet. \\item\t% The transverse motions slow down with time and cease near the time when the material reaches its maximum height and starts to fall back. The falling material travels along almost straight lines in the original direction of ascent, showing no more signatures of transverse motions. \\item\t% These observations are consistent with the scenario that the jet involves unwinding of left-handed helical threads that rotate counter-clockwise about a common axis. The untwisting wave front propagates upward at the Alfv\\'{e}n speed. The pitch spacing between adjacent helical threads increases with height, consistent with the expected axial expansion of\t% unwinding screws. The jet results in\t% magnetic helicity being shed into the upper atmosphere. \\end{enumerate} A more in-depth multiwavelength study of this event and comparison with theoretical models are underway and will be published in the future." }, "0910/0910.5715_arXiv.txt": { "abstract": "{Clusters of galaxies are believed to be capable to accelerate protons at accretion shocks to energies exceeding $10^{18}$ eV. At these energies, the losses caused by interactions of cosmic rays with photons of the Cosmic Microwave Background Radiation (CMBR) become effective and determine the maximum energy of protons and the shape of the energy spectrum in the cutoff region.} {The aim of this work is the study of the formation of the energy spectrum of accelerated protons at accretion shocks of galaxy clusters and of the characteristics of their broad band emission.} {The proton energy distribution is calculated self-consistently via a time-dependent numerical treatment of the shock acceleration process which takes into account the proton energy losses due to interactions with the CMBR.} {We calculate the energy distribution of accelerated protons, as well as the flux of broad-band emission produced by secondary electrons and positrons via synchrotron and inverse Compton scattering processes. We find that the downstream and upstream regions contribute almost at the same level to the emission. For the typical parameters characterising galaxy clusters, the synchrotron and IC peaks in the spectral energy distributions appear at comparable flux levels.} {For an efficient acceleration, the expected emission components in the X-ray and gamma-ray band are close to the detection threshold of current generation instruments, and will be possibly detected with the future generation of detectors.} ", "introduction": "\\label{Introduction} Rich clusters of galaxies are the largest virialised structures in the Universe, with typical sizes of a few Mpc and masses up to $10^{15} M_{\\odot}$ or more (see \\citealp{sarazinbook} for a review). In the standard picture of cosmic structure formation, the structure's growth is driven by gravitational instability. This process is hierarchical, with larger systems forming later via the assembly of pre-existing smaller structures. Within this scenario, galaxy clusters form via mergers, and their age can be estimated to be of the order of 10 Gyr (e.g. \\citealp{agecl}). In addition, cold material from the surrounding environment is continuously infalling, due to gravitational attraction, and an expanding shock wave, called the accretion shock, is expected to form at the cluster boundary and to carry outward the information of virialisation \\citep{bertschinger}. Numerical simulations have confirmed the appearance of the so--called accretion shocks during structure formation (e.g. \\citealp{kangsim}). The detection of a tenuous and diffuse synchrotron radio emission from about one third of rich clusters of galaxies (e.g. \\citealp{govoni}) reveals the presence of a diffuse magnetic field and a population of high energy electrons. Using exclusively synchrotron data, though, it is possible to get information only on the product of particles density and magnetic field energy density (unless adopting assumptions like equipartition). Detection of non-thermal X-rays have also been claimed from a few of such sources and generally interpreted as inverse Compton emission from the same population of electrons (e.g. \\citealp{ff, eckert}). Since IC depends on the electron density but not on the magnetic field, combining the data allows us to break the degeneracy and thus determine values of B of the order of $0.1 \\mu$G. Significantly higher values, of the order of a few microGauss, are obtained from Faraday rotation measurements \\citep{Clarke,carilli}. Thus the uncertainty in the determination of the magnetic field is quite large, allowing values to vary of about one order of magnitude \\citep{nnrephaeli}. According to theoretical models, the electrons responsible for the radio and non-thermal X-ray emission, are produced through different acceleration mechanisms (e.g. \\citealp{schlickeiser,tribble,sarazin,brunetti2001,petrosian2001}). Alternatively, the radio synchrotron radiation can be emitted by secondary electrons produced at interactions of accelerated protons with the intracluster gas \\citep{dennison,blasicol,atoyan,dolag,kushnir}. In this scenario, gamma ray emission is expected due to the decay of neutral pions produced in \\textit{p-p} interactions \\citep{volk_cluster,ber_cluster}. The non-thermal X-rays emission also can be related to synchrotron radiation of secondary electrons, but of much higher energies, produced in photon-photon \\citep{timokhin} and proton-photon \\citep{FA02,susumu} interactions. Several particle acceleration mechanisms have been proposed to operate in clusters of galaxies (see \\citealp{revste} for a review). In particular, it has been argued that large scale shocks can effectively accelerate electrons and protons up to ultrarelativistic energies \\citep{norman,ber_cluster,loeb,miniati,blasimerger,steshocks,ryu,berrington,pfrommer}. In a recent paper we performed detailed calculations of diffusive shock acceleration of electrons in galaxy clusters \\citep{mio}. Electrons can be accelerated up to 100 TeV at cluster accretion shocks, with synchrotron X-ray and IC gamma-ray radiation components produced mainly in the downstream region. While the maximum energy of electrons is limited by synchrotron and IC energy losses, the protons can be accelerated to much higher energies. Indeed, according to the Hillas criterion \\citep{hillas}, galaxy clusters are amongst the few source populations capable, as long as this concerns the dimensions of the structure and the value of the magnetic field, to accelerate protons up to $10^{20}$ eV. Moreover, clusters are cosmological structures and their lifetime is comparable to the age of the Universe, therefore, if acceleration takes place in such objects, it can continue up to $\\sim 10^{10}$ yr. On the other hand, high energy protons up to energies of the order of $10^{15}$ eV and possibly higher are well confined in the volume of the cluster over this time scale \\citep{volk_cluster,ber_cluster}. This results in an effective accumulation of high energy particles in the cluster. The thermal energy budget of rich clusters, estimated from measurements of their thermal X-ray emission, is of the order of $10^{63}-10^{64}$ erg. The non-thermal component seems to be constrained to a few percents of the thermal energy $\\sim 10^{62}$ erg \\citep{veritas,coma,magic}. This value is also comparable with the magnetic field energy, assuming 1 $\\mu$G and a spherical cluster of 3 Mpc of radius (e.g. the size of the Coma cluster). Although the dimensions of the system, the strength of the magnetic field, and the age of the accelerator formally allow protons to be accelerated up to $10^{20}$ eV, the particles lose, in fact, their energy via pair and pion production in the interactions with the photons of the Cosmic Microwave Background (CMBR) radiation field. As discussed in \\citet{norman} and \\citet{kang_rach_bier}, for a shock velocity of a few thousands km/s and a magnetic field of the order of $1 \\mu \\rm G$, the shock acceleration rate is compensated by the energy loss rate at energies around $10^{19}$ eV (the exact value depends on the assumed diffusion coefficient and also on the shock geometry as shown by \\citealp{ostrowski}). The interactions of ultra-high energy protons with the CMBR lead to production of electrons in the energy domain ($\\geq 10^{15} \\ \\rm eV$) which is not accessible through any direct acceleration mechanism. These electrons cool via synchrotron radiation and IC scattering on very short timescales (compared to both the age of the source and interaction timescales of protons). For the same reason they are rather localised in space, i.e. are ``burned\" not far from the sites of their production. Therefore the corresponding radiation components in the X-ray and gamma-ray energy bands are precise tracers of primary protons, containing information about the acceleration and propagation of their ``grandparents\". This interesting feature has been indicated in \\citet{FA02} in the general context of acceleration and propagation of ultrahigh energy protons in large scale extragalactic structures. More specifically, this issue was discussed by \\citet{susumu} for objects like the Coma galaxy cluster. However in these studies an {\\it a priori} spectrum of protons has been assumed in the ``standard form\" $E^{-2} \\exp(-E/E^*)$. In fact, the energy distribution of protons accelerated in galaxy clusters via Diffusive Shock Acceleration (DSA) mechanism can be quite different from this simple form, as it is demonstrated below. In this paper we study the process of proton acceleration by accretion shocks in galaxy clusters taking into account self-consistently the energy loss channels of protons related to their interactions with the CMBR photons. We make use of the numerical approach presented in \\citet{mio}. In Section \\ref{Proton Acceleration and Energy losses} the calculation is introduced and the accelerated proton spectra are derived. In particular, we show that electron/positron pair production is the dominant energy loss channel for protons. In Section \\ref{Spectra of Secondary Electrons} we calculate the spectra of secondary pairs. The broad-band emission produced by the secondary electrons during their interactions with the background magnetic and radiation fields is presented in Section \\ref{Radiation Spectra}. In Section \\ref{Shock Modification} we study, in a simplified fashion, the impact of shock modification by efficiently accelerated protons on the acceleration and emission features. We briefly discuss and summarise the main results in Section \\ref{Conclusions}. ", "conclusions": "\\label{Conclusions} Proton acceleration in galaxy clusters was studied in the framework of DSA via a detailed time-dependent numerical calculation that includes energy losses due to interactions of protons with photons of the CMBR. For realistic shock speeds of a few thousand km/s and a background magnetic field close to $1 \\mu \\rm G$, the maximum energy achievable by protons is determined by the energy losses due to pair production and ranges from a few times $10^{18}$ eV to a few times $10^{19}$ eV. We performed the calculations assuming that acceleration takes place on time scales comparable to the age of the cluster. Since steady state is never achieved in this scenario, a time-dependent treatment is required. Particle spectra, when calculated including the effect of energy losses, exhibit interesting features. The decay of the spectrum above the cut-off energy is not exponential. Its dependence on energy is shallower due to the flat profile of pair production timescales in the cut-off energy range. The time-dependent distributions of protons are used to calculate accurately the production rates of secondary electron-positron pairs. These electrons cool rapidly via synchrotron radiation and IC scattering which proceeds in the Klein-Nishina regime. The effect of the hardening induced by the KN cross-section is visible in the IC radiation spectra both in the upstream and downstream regions of the shock. For the fiducial Coma-like cluster used in this work, the synchrotron and IC peaks of the electron broadband SED are at comparable levels and the associated flux from a source at the distance of a 100 Mpc is expected at the level of $10^{-12}$ erg cm$^{-2}$ s$^{-1}$ in the X-rays and an order of magnitude lower for TeV gamma-rays. Note that although the maximum of the gamma-ray emission is located above 100 TeV, it unfortunately cannot be observed due to severe intergalactic absorption. The expected gamma-ray flux from clusters of galaxies is at the limit of the sensitivity of present generation instruments, however it may be detectable with the future generation of detectors. The optimum energy interval for gamma-ray detection is between 1 and 10 TeV. The detectability of clusters in hard X-rays and gamma-rays associated with interactions of ultrahigh energy protons with the CMBR, depends on the value of the parameter $A=W_{62}/d_{100}^2$, where $W_{62}=W/10^{62} \\ \\rm erg$ is the total energy released in cosmic rays normalised to $10^{62}$ erg, and $d_{100}=d/100$ Mpc is the distance normalised to 100 Mpc. Obviously, the best candidates for detection are nearby rich galaxy clusters like Coma and Perseus located at distances $d \\sim 100$ Mpc. On the other hand, due to the large extension of the non-thermal emission (as large as several Mpc), and given that the angular size of the source $\\theta \\propto 1/d$, the probability of detection reduces with the distance slower than $1/d^2$. Nevertheless, as long as the total energy in accelerated protons does not significantly exceed $10^{62}$ erg, the visibility of clusters of galaxies in X-rays and gamma-rays is limited by objects located within a few 100 Mpc. The chances of detection of non-thermal emission of clusters related to ultrahigh energy protons, especially in the hard X-ray band, can be significantly higher if protons are accelerated by non-linear shocks modified by the pressure of relativistic particles. In this scenario, a large fraction of the energy of the shock is transferred to relativistic protons. Moreover, in this case a very hard spectrum of protons is formed, thus the main fraction of non-thermal energy is carried by the highest energy particles. These two factors can enhance the luminosity of X- and gamma-ray emission of secondary electrons by more than an order of magnitude, and thus increase the probability of detection of clusters located beyond 100 Mpc. In this paper we do not discuss the gamma-ray production related to interactions of accelerated protons with the ambient gas which can compete with the inverse Compton radiation of pair produced electrons. The relative contributions of these two channels depends on the density of the ambient gas and the spectral shape of accelerated protons. The flux of gamma-rays from {\\it pp} interactions can be easily estimated based on the cooling time of protons, $t_{\\rm pp} \\approx 1.5 \\times 10^{19}/n_{-4}$ s, where $n_{-4}=n/10^{-4}$ cm$^{-3}$ is the density of the ambient hydrogen gas, normalised to $10^{-4}$ cm$^{-3}$. Then, the energy flux of gamma-rays at 1 TeV is estimated as $F_\\gamma (\\sim 1 \\rm TeV) \\approx 6 \\times 10^{-12} \\kappa W_{62} n_{-4}/d_{100}$ erg cm$^{-2}$ s$^{-1}$, where $\\kappa$ is the fraction of the total energy of accelerated protons in the energy interval between 10 to 100 TeV (these protons are primarily responsible for production of gamma-rays of energy $\\sim 1$ TeV). For a proton energy spectrum extending to $10^{18}$ eV, this fraction is of order of $\\kappa \\sim 0.1$. Thus for an average gas density in a cluster like Coma, $n \\sim 3 \\times 10^{-4}$ cm$^{-3}$, the gamma-ray flux at 1 TeV is expected at the level of $10^{-12}$ erg cm$^{-2}$ s$^{-1}$ which is comparable to the contribution of IC radiation of secondary electrons. In the case of harder spectra of protons accelerated by non-linear shocks, the contribution of gamma-rays from {\\it pp} interactions is dramatically reduced and the contribution of secondary pairs to gamma-rays via IC scattering strongly dominates over gamma-rays from {\\it pp} interactions." }, "0910/0910.1136_arXiv.txt": { "abstract": "{OH maser emission at 1.67 GHz is known to be associated with regions of intense star formation within ULIRGs. As these galaxies are formed through violent mergers, studying the co-moving density of OH maser galaxies across cosmic time will allow the merger rate of the Universe to be determined in an independent way. This merger rate is an important parameter in galaxy formation and evolution scenarios. The sensitivity, wide field of view and spectral coverage of both APERTIF on the WSRT and ASKAP will allow for the first time all-sky blind surveys for OH maser galaxies to be carried out to redshift 1.4. We describe the prospects for such surveys, including the expected number of OH maser galaxies that will be discovered, and what limits can be placed on the OH maser luminosity function, and hence the merger rate out to redshift 1.4 with various survey strategies.} \\FullConference{Panoramic Radio Astronomy: Wide-field 1-2 GHz research on galaxy evolution - PRA2009\\\\ June 02 - 05 2009\\\\ Groningen, the Netherlands} \\begin{document} ", "introduction": "The 1.665 GHz and 1.667 GHz doublet transitions of hydroxyl (OH) show spectacularly luminous maser emission in regions of intense circumnuclear star-formation, with reported line luminosities of up to $10^4~L_{\\odot}$ (for a review see \\cite{lo05}; for an example of the OH doublet maser lines see Figure \\ref{maser-red}). The masing action results from a strong far-infrared radiation field, generated by the heated dust associated with star-formation, that pumps the OH molecules into an excited state \\cite{baan85,baan89}. The most powerful OH masers are found in the population of ultra-luminous infrared galaxies (ULIRGs; $L_{FIR} \\geq 10^{12}~L_{\\odot}$) that are believed to form through galaxy mergers. Indeed, surveys of ULIRGs at low redshifts have found up to one third to show OH maser emission \\cite{darling02a}, with a correlation found between the far-infrared and OH maser luminosities ($L_{OH} \\propto L_{FIR}$$^{2.3}$; \\cite{kloeckner04}). Thus, one of the most powerful applications of OH masers is their unbiased ability to trace the number density of ULIRGs and hence the galaxy merger history of the Universe, which forms a fundamental part of galaxy formation and evolution scenarios (e.g. \\cite{croton06}). There are to date $\\sim$100 OH maser galaxies known and all are found in the local Universe ($z \\leq 0.3$; see Figure \\ref{maser-red} for the OH maser redshift distribution). By detecting OH masers at higher redshifts, it will be possible to determine the density evolution of the OH maser luminosity function, and hence make an independent measurement of the merger rate of the Universe as a function of redshift. In the future, wide-field surveys with the next generation of radio telescopes and arrays can be carried out to determine the evolution in the OH maser luminosity function. The potential for using OH maser galaxies in this way was pointed out by Briggs \\cite{briggs98}. In this paper, we discuss the prospects for such surveys with two of these new facilities; the first is APERTIF (APERture Tile In Focus) on the Westerbork Synthesis Radio Telescope and the second is ASKAP (the Australian Square Kilometre Array Pathfinder). \\begin{figure}[tbh] \\begin{center} \\includegraphics[width=\\textwidth]{maser-red.eps} \\caption{Left: an example of the 1.665 GHz and 1.667 GHz doublet transitions of hydroxyl from the galaxy 10035+2740 at redshift 0.1662, taken from \\cite{darling02a}. Right: the redshift distribution of extragalactic OH maser galaxies. Note that this redshift distribution is not complete, but is based on targeted observations of candidate OH maser galaxies, for example, ULIRGs.} \\label{maser-red} \\end{center} \\end{figure} ", "conclusions": "\\label{conc} We have presented new simulations of the potential numbers of OH maser galaxies that can be detected with APERTIF and ASKAP out to redshift $\\sim$1.4. Our results suggest that with 3 months of observing time, it will be possible to cleanly differentiate between evolutionary models for the OH maser population. Given that these new facilities will be carrying out large all-sky blind surveys for H\\,{\\sc i} in the same frequency range that we would expect to find high redshift OH, it would require only a small additional amount of effort to expand the search to look for OH galaxies. The largest uncertainty in our analysis comes from the poorly constrained local OH maser luminosity function at high luminosities ($\\geq 10^{3.8}~L_{\\odot}$). There are a number of avenues available to try and constrain the high-luminosity end of the luminosity function in the short term, which would further constrain the predictions made here. The first is to carry out blind searches of small fields with current instruments, for example, with the Giant Metrewave Radio Telescope, as suggested by Darling \\& Giovanelli \\cite{darling02b}. Another option is to observe gravitationally lensed ULIRGs to detect OH masers at even higher redshifts ($z>2.5$). Here, the lensing magnification can be used to detect faint lines which would otherwise not be observable without the magnification of the lens. This technique was recently used to detect the most distant water maser from a high redshift galaxy \\cite{impellizzeri08}." }, "0910/0910.4978_arXiv.txt": { "abstract": "{} {It is widely recognized that processes taking place inside group environment could be among the main drivers of galaxy evolution. Compact groups of galaxies are in particular good laboratories for studying galaxy interactions and their effects on the evolution of galaxies due to their high density and low velocity dispersion. SCG0018-4854 is a remarkably high galaxy density and low velocity dispersion group with evidence of a recent interaction. It has been detected and analyzed at different wavelengths, but its kinematics has not yet been studied in detail.} {We obtained VLT FORS2 optical observations and we present spectroscopic and photometric evidence of how dramatically galaxy interactions have affected each of the four member galaxies.} {We found peculiar kinematics for each galaxy and evidence of recent star formation. In particular, the gas and stellar radial velocity curves of two galaxies are irregular with a level of asymmetry similar to that of other interacting galaxies. We discovered the presence of a bar for NGC 92 therefore revising a previous morphological classification and we obtained spectroscopic confirmation of a galactic-scale outflow of NGC 89.} {Peculiar kinematics and dynamic consideration lead to a rough estimate of the age of the latest interaction: $\\tau \\approx 0.2-0.7 \\, \\mathrm{Gyr}$, suggesting that SCG0018-4854 is a young and dynamical group.} ", "introduction": "Compact groups (CGs) are among of the smallest and densest systems of galaxies of the universe, furthermore they are characterized by low velocity dispersions of the order of $\\approx 200\\, \\mathrm{km \\, s^{-1}}$ \\citep{1992ApJ...399..353H}. This is why they are excellent laboratories for studying galaxy interactions and their effects on the evolution of galaxies. \\citet{1991ApJS...76..153R} showed that 30\\% of spiral galaxies in CGs has distorted rotation curves and another 30\\% has irregular velocity patterns indicating that tidal interactions are frequent and persisting in CGs. Moreover, \\citet{1994ApJ...427..684M} demonstrated that 43\\% of the galaxies in their compact group sample shows morphological and/or kinematical distortions indicative of interactions and/or mergers. Since these pioneering studies it has been clear that interactions play an important role in the evolution of groups. According to current dynamical models, galaxies belonging to groups should interact violently and merge on a relatively short time scale \\citep{1987ApJ...322..605S,1985MNRAS.215..517B,1996ApJ...458...18G}. Yet, strong mergers in CGs are rare \\citep{1994ApJ...427..684M} and a lot of effort has been made to establish the origin and fate of such systems. Recent studies \\citep{2007AJ....133.2630C} have shown that galaxies in compact groups are more likely to merge under dry conditions, after they have lost most of their gas in interactions among galaxies. These hypothesis is corroborated by HI studies; \\citet{2001A&A...377..812V} have found out perturbed distributions of neutral hydrogen in CGs. This results indicate that the evolution of CGs is mainly dominated by galaxy interactions, continuous tidal stripping and/or gas heating. It is therefore interesting to reconstruct the evolution and the interaction history of a CG throughout the properties of its galaxy members and compare them with group properties as a whole. Kinematical and dynamical information are fundamental to separate between different interaction histories. \\citet{1998ApJ...507..691M} defined some kinematic indicators to distinguish between merging and interaction, such as misalignment between the kinematic and photometric axis of gas and stellar components, double gas components, warping and other peculiarities. However, there are still open questions about how many interactions take place and how strong and frequent they are. In this work we focus on SCG0018-4854, a spiral-only compact group characterized by a really dense environment in which member galaxies have clearly interacted. The plan for the article is the following. In Sect.~\\ref{sec:SCG0018-4854}, we present the characteristics of SCG0018-4854. In Sect.~\\ref{sec:obs}, we describe our observations and the reduction process. In Sect.~\\ref{sec:method}, we explain the methods used for our analysis. The results are presented in Sect.~\\ref{sec:results}, followed, in Sect.~\\ref{sec:Discussion}, by a short discussion and our conclusions in Sect.~\\ref{sec:Conclusions}. ", "conclusions": "\\label{sec:Conclusions} We have analyzed high signal-to-noise spectra along both the major and the minor axis of each galaxy of SCG0018-4854. Each galaxy shows disturbed or peculiar kinematics. Kinematic information is important in order to have some insight about the possible interaction history of the galaxies. Different interaction scenarios, depending on the strength and the geometry of the encounter and the morphological types of the interacting systems, will leave different signatures on the galaxy kinematics. Following \\citet{1998ApJ...507..691M} we found out signs of disturbed velocity fields, asymmetric rotation curves, multiple kinematic gas components, tidal tails and nuclear activity. All these characteristics suggest that at least three out of four galaxies of SCG0018-4854 have been involved in a strong and recent interaction. We derived some constraints for the age of the latest close interaction: $\\approx 190 \\, \\mathrm{Myr} < \\mathit{\\tau_{coll}} < 0.7 \\, \\mathrm{Gyr}$. The interaction has been strong enough to perturb the gas kinematics up to a level of 24\\% for the gas of the two main galaxies. These results are in agreement with the recent claim by \\citet{2009arXiv0908.2798T} that this group is in an advanced stage of interaction, based on the presence of young star forming regions and a candidate tidal dwarf galaxy identified in the UV in addition to the observed tidal tails. Finally, the results of our spectroscpic analysis are consistent with the hypothesis of a large-scale outflow in NGC 89. In contrast to its evolved properties as a group such as the high spatial density, the spatial distribution of HI and the presence of a common hot IGM, SCG0018-4854 is entirely composed of spiral galaxies and has a remarkably high fraction of active galaxies. Moreover, this study reveals kinematical signs of recent interactions among its members making SCG0018-4854 a dynamically young group although its global properties suggest a more advanced stage of evolution. Given the estimated age of the latest interaction, we could say that we catch these galaxies in the act of interacting. What about the future of this group? Hydrodynamical simulations could help us to understand the fate of SCG 0018-4854, a really isolated interacting compact group." }, "0910/0910.1593_arXiv.txt": { "abstract": "Merging compact binaries are the one source of gravitational radiation so far identified. Because short-period systems which will merge in less than a Hubble time have already been observed as binary pulsars, they are important both as gravitational wave sources for observatories such as LIGO but also as progenitors for short gamma-ray bursts (SGRBs). The fact that these systems must have large systemic velocities implies that by the time they merge, they will be far from their formation site. The locations of merging sites depend sensitively on the gravitational potential of the galaxy host, which until now has been assumed to be static. Here we refine such calculations to incorporate the temporal evolution of the host's gravitational potential as well as that of its nearby neighbors using cosmological simulations of structure formation. This results in merger site distributions that are more diffusively distributed with respect to their putative hosts, with locations extending out to distances of a few Mpc for lighter halos. The degree of mixing between neighboring compact binary populations computed in this way is severely enhanced in environments with a high number density of galaxies. We find that SGRB redshift estimates based solely on the nearest galaxy in projection can be very inaccurate, if progenitor systems inhere large systematic kicks at birth. ", "introduction": "The association of short gamma-ray bursts with both star-forming galaxies and with ellipticals dominated by old stellar populations \\citep{Berger05,Bloom05,Fox05,Gehrels05,Prochaska05,Berger09} suggested an analogy to type Ia supernovae, as it indicated a class of progenitors with a wide distribution of delay times between formation and explosion. Similarly, just as core-collapse supernovae are discovered almost exclusively in star-forming galaxies, so too are long GRBs \\citep{wbrev}. Indeed, a detailed census of the types of host galaxies, burst locations and redshifts could help decide between the various SGRB progenitor alternatives \\citep[e.g.][]{zr07,guetta05,bp06}: If the progenitor lifetime is long and the systemic kick is small, then the bursts should correspond spatially to the oldest populations in a given host galaxy. For early-type galaxies, the distribution would most likely follow the light of the host. A neutron star (NS) binary could take billions of years to spiral together, and could by then, if given a substantial kick velocity on formation, merge far from its birth site \\citep{fryer99,Bloom99,bbk02,Bloom02}. The burst offsets would then most likely be larger for smaller mass hosts. Double neutron star binaries, such as the famous PSR1913+16, will eventually coalesce, when gravitational radiation drives them together \\citep{vk07}. Each supernova is thought to impart a substantial kick to the resulting NS \\citep{hp97}. For systems that survive both supernovae explosions the center of mass of the remnant binary itself will receive a velocity boost on the order of a few hundred kilometers per second \\citep{bp95,fk97}. As a result, NS binaries will be ejected from their birth sites. The exact distribution of merger sites depends sensitively on the gravitational potential of the host, which until now has been assumed to be static. The potential of a realistic host galaxy is, however, not static. In fact, the gravitational potential of the host as well as that of its nearby neighbors is expected to evolve dramatically from compact binary production until coalescence. In order to incorporate these effects self-consistently, in this {\\it Letter}, we study the orbital evolution of compact binary systems using cosmological simulations of structure formation. Our results provide new insights into what happens when compact binary stars are ejected from their birth halos as a result of velocity kicks, and what progenitor clues a distant observer might uncover from the distribution of SGRB sites in and around galaxies. ", "conclusions": "Before the detailed modeling of light curves was used to constrain the nature of supernovae progenitors, the location of supernovae in and around galaxies provided important clues to the nature of the progenitors. Similarly, in the absence of supernova-like features \\citep[e.g.][]{hjorth05a}, detailed observations of the astrophysics of individual host galaxies may thus be essential before stringent constraints on the identity of SGRB progenitors can be placed\\footnote{The interaction of burst ejecta with a stellar binary companion \\citep{andrew} or with its emitted radiation \\citep{enrico} could also help shedding light on the identity of the progenitor system.} \\citep[e.g.][]{zr07,Fox05}. Even with a handful of SGRBs detected to date, it has become apparent that short and long events are not drawn from the same parent stellar population \\citep{nakar}. In contrast to long GRBs, the galaxies associated with SGRBs exhibit a wide range of star-formation rates, morphologies and metallicities \\citep{Berger09}. They are also often found in older and lower-redshift galaxies and, in a few cases, with large ($\\gtrsim$ 10 kpc) projected offsets from the centers of their putative host galaxies \\citep{Bloom05,Prochaska05,Berger05,Gorosabel06,levan,bloom07}. SGRBs are, however, not universally at large offsets and are not always associated with early-type galaxies \\citep{Fox05,villa,hjorth05b,covino,as06}. The discovery of early-type galaxy hosts suggest a progenitor lifetime distribution extending well beyond a Gyr. A large progenitor lifetime would help explain the apparent high incidence of galaxy cluster membership \\citep{Bloom05,Ped05,Berger07a}. On the other hand, shorter lifetimes are required to explain the population of SGRB at moderately high redshift \\citep{Berger07b,graham09}. The observed projected distances from what has been argued are the plausible hosts, if true, also holds important ramifications for the sort of viable progenitors \\citep{bp06}. The large offsets seen from early-type hosts would seem to be at odds with progenitor systems with small systematic kicks \\citep[such as in globular clusters][]{Grindlay06,lrr07,rion}, although with such large physical offsets the possibility remains that the association with the putative host is coincidental. On the other hand, based on the small offsets from some low-mass galaxy hosts \\citep{Prochaska05,as06,Berger09,bp06}, SGRB progenitors cannot all have large systematic kicks at birth and inhere large delay times from formation. Making more quantitative statements about the nature of the progenitor systems is not only hampered by small number statistics but also from the lack of robust predictions of the distribution of merger sites. These distributions depend sensitively on the gravitational potential of the host, which until now has been assumed to be static, and the compact binary formation properties, especially the systematic kick velocity. Here we have refined such calculations to incorporate the temporal evolution of the host's gravitational potential as well as that of its nearby neighbors self-consistently using cosmological simulations of structure formation. This results in diverse predictions of offsets and compact binary demographics even in the simplest case of a kick velocity distribution \\citep[here assumed to be in excess of $200\\;{\\rm km\\;s^{-1}}$ in order to explain the observed parameters of double neutron star systems;][]{fk97} whose properties do not vary with the initial binary separation. Two important predictions stand out. First, the merger site distributions computed in this way are more diffusively located with respect to their putative hosts. In a field environment, for example, the distribution of merging sites can extend out to a distance of a few Mpc. This is more severe for those host galaxy halos that were unable to effectively retain most of its own compact binary population at birth. Second, the degree of mixing between neighboring compact binary populations depends on galactic environment. In a cluster, for example, the mixing of the various compact binary populations is severe as a result of the high merger activity and increases dramatically with time. At $z=0$, in a cluster environment, the probability of finding a foreign coalescing compact binary system is equal or higher at all radial distances than finding one originating from the massive central halo. As a result, the closest galaxy at the time of binary coalescence and possibly SGRB occurrence may not to be the one where the compact binary system originated from. Of course, our basic model is rather simple since we assume a single epoch of star formation and a simple star formation recipe. Also different distributions of kick velocities should be considered. We do not expect that our qualitative results will change dramatically as our first, more general, results indicate. It is evident form the discussion above that assuming a large (already evolved) host galaxy at the time of compact binary formation thus severely overestimates the binary retention fraction and the concentration of their merging site distribution. This implies that SGRB redshift estimates based solely on the nearest galaxy in projection can be very inaccurate, if progenitor systems inhere large systematic kicks at birth. Interpretations on the nature of the SGRB progenitor using the stellar and mass properties of the nearest galaxy in projection as established by the afterglow location must therefore be regarded with suspicion. Finally, it should be noted that a direct comparison with model predictions is still impeded by the possibility of an ambient density bias \\citep{Bloom05,Lee05} where SGRBs are more likely to be found in denser gas regions and , as a result, we could be missing a population of bursts with large systematic kicks." }, "0910/0910.2379_arXiv.txt": { "abstract": "A method is proposed for constraining the Galactic gravitational potential from high precision observations of the phase space coordinates of a system of relaxed tracers. The method relies on an ``ergodic'' assumption that the observations are representative of the state of the system at any other time. The observed coordinates serve as initial conditions for moving the tracers forward in time in an assumed model for the gravitational field. The validity of the model is assessed by the statistical equivalence between the observations and the distribution of tracers at randomly selected times. The applicability of this ergodic method is not restricted by any assumption on the form or symmetry of the potential. However, it requires high recision observations as those that will be obtained from missions like SIM and GAIA. ", "introduction": "Measurements of velocities of cosmological objects are a classic probe of the mass distribution on all scales. It was the motions of individual galaxies in galaxy clusters which first showed that luminous mater contributed only a small fraction to the total mass in clusters \\cite[]{Zwicky37}, implying the existence of dark matter. The combined mass of the Milky Way (or the Galaxy) and M31 is constrained from the observed relative motion between the two galaxies and the requirement that the initial distance between their respective centers of mass vanishes near the Big Bang \\cite[]{KW59}. Line-of-sight velocities of other galaxies in the Local Group of galaxies also are used to estimate its mass by the condition of vanishing initial distances \\cite[]{Peebles89}. On scales 10s of Mpcs, peculiar motions (deviations of Hubble flow) of galaxies constrain the global mass density in the Universe \\cite[e.g.][]{Nusser08}. Determining the mass distribution in the Milky Way is particularly important. There is ample information on the baryonic content of the Galaxy which could be modeled in detail only if the dark matter distribution is known. The rotation curve of the Galaxy is limited to distances smaller than 20 kpc and does not provide any information on deviations from spherical symmetry of the halo. The mass distribution at larger distances, motions of Galactic satellite galaxies, globular clusters and stars are invoked \\cite[e.g.][]{sakamoto03}. Constraints on the Galactic mass from these tracers are derived from the condition that the observed speeds of Galactic objects do not exceed the escape velocity. This approach yields only a lower limit and is mostly sensitive to the highest velocity objects. The vast majority of the sample objects play no role in deriving the mass limit. Alternatively, one could adopt a Bayesian likelihood formalism in which the phase space distribution function is assumed to follow certain form which could be matched with the observations to probe the Galactic potential field \\cite[e.g.][]{LT87,Koch96,WE99} . Proper motions, resulting from velocities perpendicular to the line-of-sight, are currently measured only for nearby tracers \\cite[e.g.][]{sakamoto03}. Therefore, most current mass estimates rely on the measured line-of-sight motions of individual components. During the next few years, accurate proper motions for a large sample of Galactic tracers are expected to be measured by the space missions {\\it Global Astrometry Interferometer for Astrophysics } (GAIA) \\cite[]{LP96} and {\\it Space Interferometry Mission } (SIM) \\cite[]{rr1}. Even accurate phase space information require additional assumptions in order to constrain the Gravitational potential \\footnote{The gravitational force field is equal to the acceleration rather than velocities. Measuring the acceleration of a tracer with orbital period $\\td$ over an observing time $\\tobs$ requires astrometry with angular resolution a factor $\\tobs/\\td$ higher than the precision needed for velocities. Since $\\tobs\\sim$ a few years while $\\td\\sim$ a few Gyrs, the task is out of reach in the near future. }. We present here a general method which relies on high precision measurements of positions and velocities of tracers. The method assumes that the tracers have reached an equilibrium state in the Galactic gravitational field. ", "conclusions": "\\label{sec:Conclude} The ergodic method presented here requires a parametric functional form for the Gravitational potential, but does not impose any special symmetry on the mass distribution. The method assumes that the observations of a class of tracers are spatially complete. Tracers with observed distances smaller than $d_0$, could be present beyond $d_0$ in snapshots at later times. Therefore, Observational selection against tracers at distances $>d_0$ will make the statistical comparison between observations and snapshots at later times extremely difficult. The completeness is not too demanding a condition for tracers like globular clusters and Galactic satellites for which future observations should be accurate enough for quite large Galactic distances. We have presented only partial testing of the method, with only a one parameter family for the form of the Galactic potential. Method is able to constrain the parameter to a good accuracy with measurements errors that are even larger than those expected to be achieved by future data. The tests show that the method could provide unbiased constraints on the Gravitational potential, but a more elaborate testing which includes a more realistic treatment of the errors should be done. Observations will likely assign distance and velocity measurement errors to tracers on an individual basis. Therefore, random errors and systematic biases tailored to the specific sample of tracers used by the method could be determined robustly. The accuracy of the method is mainly limited by the number of tracers. The most obvious tracers are globular clusters and Galactic satellites. Our Galaxy includes 158 known globular clusters, and 23 known satellites \\cite[e.g.][]{SG07}. Distance measurements of RR Lyrae stars from their period-luminosity relation will be greatly improved by GAIA and SIM calibration of the zero point using a nearby sample of these stars. Therefore, luminous halo RR Lyrae stars could significantly enlarge the sample of tracers. The method requires a system of tracers in dynamical equilibrium in the current Galactic potential. A pre-requist for dynamical equilibrium is that any recent changes in the Galactic potential must have occurred on a time scale longer that the dynamical time of the system of tracers. Spectroscopic studies stars in the Galaxy do not present evidence for substantial mergers in the last 8Gyr \\cite[]{GWN,Helmi}, implying a nearly static gravitational potential. We have demonstrated that adiabatic growth of the Galactic disk is not expected to pose a problem for the implementation of the method. The effect of any deviation from that should be modeled. One issue which is exclusive to using disk stars as tracers is the effect of transient perturbations on the dynamics of those stars \\cite[e.g.][]{FAM05,QUIL}. However, the transient effects cause velocity perturbations at the level of ~10 km/s which is much smaller that the total velocities of tracers. Observational uncertainties are larger than that and do not seem to cause significant biases in the results as indicated by the tests described above. Another issue to be considered is halo substructure which, in principle, could act as a stochastic component in the gravitational potential. Such a component is extremely difficult to model in the method proposed here. We offer the following argument demonstrating that substructure should not have an important effect on the long term dynamics of tracers. In the limit of fast encounters, a tracer passing a substructure at distance $b$ will change its velocity by $V_1\\approx g(b)b/V$ where $V$ is the relative velocity and $g\\approx G M/b^2$ is the gravitational force field of the substructure assuming that $b$ is larger than its tidal radius. Performing the usual summing in quadratures over encounters occurring in one orbital time we get that the r.m.s change $^{1/2}\\approx M\\sqrt{\\bar n R}/V$ where $\\bar n$ is the number density of substructures and $R$ is the distance travelled by the tracer in one orbital time. Taking $V^2=G M_{_{\\rm MW}}(R)/R$ with $M_{_{\\rm MW}}(R)$ the mass of the Galactic halo within radius $R$, we get the condition $\\sim M_{_{\\rm MW}} \\sqrt{\\bar n R^3}>M_{_{\\rm MW}}$ for having $^{1/2}\\sim V$. This means that neither single encounters nor collective stochastic effects can dominate the long term evolution of tracers even if the fraction of mass in substructures is large which is contrary to recent simulations \\cite[e.g.][]{Colombi} which show that the smooth component greatly dominates the mass of Galactic size halos. When details of future data from SIM and GAIA become available, all robust information about the distribution of baryons in the Galaxy should be used \\cite[e.g.][]{ROBIN03} in order to place tight constraints on the Galactic dark matter. The validity of the method should be tested with mock data that match the observations as much as possible and with the best possible available Galactic models." }, "0910/0910.0183_arXiv.txt": { "abstract": "Based on ideas by Woodward et al., a subgrid scale model that is applicable to highly compressible turbulence is presented. Applying the subgrid scale model in large eddy simulations of forced supersonic turbulence, the bottleneck effect is largely reduced and, thereby, approximate scaling laws can be obtained at relatively low numerical resolution. In agreement with previous results from PPM simulations without explicit subgrid scale model, it is found that the energy spectrum function for the velocity field with fractional density-weighing, $\\rho^{1/3}\\vect{u}$, varies substantially with the forcing, at least for the decade of wavenumbers next to the energy-containing range. Consequently, if universal scaling of compressible turbulence exists, it can be found on length scales much smaller than the forcing scale only. ", "introduction": "Large eddy simulations (LES) are of great utility to engineers and atmospheric scientists, but not commonly used in computational astrophysics. While terrestrial turbulence is incompressible or weakly compressible in most cases, astrophysicists often deal with highly compressible turbulence. Since the stability of numerical solvers for the equations of compressible gas dynamics requires energy dissipation that is intrinsic to the numerical scheme, gas-dynamical simulations at high Reynolds numbers are considered as \\emph{implicit} large eddy simulations. However, it was shown that numerical schemes such as the piecewise parabolic method (PPM) of \\citet{ColWood84} entail undesired properties of the numerical solutions such as the bottleneck effect, i.~e., an unphysical enhancement of spectral power in the high-wave number range \\citep[e.~g.,][]{KritNor07}. This led \\citet{WoodPort06} to the idea to couple an \\emph{explicit} subgrid-scale (SGS) model to PPM, and they applied the method to decaying transonic turbulence. In this article, I briefly outline an SGS model that is valid in the supersonic regime. As a first application, I present results from LES of forced supersonic isothermal turbulence with root mean square Mach number between $5$ and $6$, which can be interpreted as an idealized model for cold turbulent gas in the interstellar medium \\citep[see][]{KritNor07,FederDuv09}. Computing turbulence energy spectra from these LES, I find that the SGS model largely reduces the bottleneck effect in comparison to plain PPM simulations. The results shed light on the question of the universality of compressible turbulence, which is important for the theory of star formation in molecular clouds \\citep[see, for example,][]{ElmeSca04}. ", "conclusions": "The bottleneck effect, which distorts turbulence energy spectra toward high wavenumbers in PPM simulations without explicit SGS model, is mostly compensated by the SGS turbulence stress in LES. This property is clearly favorable for the determination of scaling laws from the spectra. Moreover, the SGS model for compressible turbulence presented in this article will be valuable in simulations of complex astrophysical flows, where a large range of scales covers energy-injecting physical processes, while only a small portion of the inertial subrange can be resolved. A first example for such an applications has recently been presented by \\citet{MaierIap09} Based on new data for hydrodynamical as well as magnetohydrodynamical turbulence, \\citet{KritUst09} put forward the hypothesis that the universality of compressible turbulence becomes manifest in the scalings of $\\rho^{1/3}\\vect{u}$. The spectral indices inferred from the $512^{3}$ LES presented in this article, however, do not support this hypothesis. There are two interpretations of these results. One interpretation presumes that the scalings in the range of wavenumbers next to the energy-containing range reflect genuine inertial-range properties of supersonic turbulence. Then the turbulence energy spectra clearly exhibit a dependence of these properties on the forcing \\citep[also see][]{SchmFeder08}. As proposed by \\citet{FederDuv09}, the prominent flattening of the spectrum function $E_{1/3}(k)$ at high wavenumbers that is observed for compressive forcing (see right plot in Fig.~\\ref{fg:spect_512}) might be caused by a transition to the weakly compressible regime at the sonic wavenumber, for which the turbulent velocity fluctuations are comparable to the speed of sound. On the other hand, it is possible that compressible turbulence with universal scaling can only develop on length scales much smaller than forcing scale if the forcing is mostly compressive. In order to settle the question which interpretation is correct, the combined influence of varying the Mach number of the flow and choosing different mixing ratios of solenoidal and compressive forcing components has to be investigated. In this regard, the SGS model presented in this article can help to explore the parameter space." }, "0910/0910.5590_arXiv.txt": { "abstract": "{We present new optical long-slit data along 6 position angles of the bulge region of M31. We derive accurate stellar and gas kinematics reaching 5 arcmin from the center, where the disk light contribution is always less than 30\\%, and out to 8 arcmin along the major axis, where the disk makes 55\\% of the total light. We show that the velocity dispersions of McElroy (1983) are severely underestimated (by up to 50 km/s). As a consequence, previous dynamical models have underestimated the stellar mass of M31's bulge by a factor 2. As a further consequence, the light-weighted velocity dispersion of the galaxy grows to 166 km/s and to 170 km/s if also rotation is taken into account, thus reducing the discrepancy between the predicted and measured mass of the black hole at the center of M31 from a factor 3 to a factor 2. The kinematic position angle varies with distance, pointing to triaxiality, but a quantitative conclusion can be reached only after simultaneous proper dynamical modeling of the bulge and disk components is performed. We detect gas counterrotation near the bulge minor axis. We measure eight emission-corrected Lick indices. They are approximately constant on circles. Using simple stellar population models we derive the age, metallicity and $\\alpha$-element overabundance profiles. Except for the region in the inner arcsecs of the galaxy, the bulge of M31 is uniformly old ($\\ge 12$ Gyr, with many best-fit ages at the model grid limit of 15 Gyr), slightly $\\alpha$-elements overabundant ($[\\alpha/Fe]\\approx 0.2$) and at solar metallicity, in agreement with studies of the resolved stellar components. The predicted u-g, g-r and r-i Sloan color profiles match reasonably well the dust-corrected observations, within the known limitations of current simple stellar population models. The stellar populations have approximately radially constant mass-to-light ratios ($M/L_R\\approx 4-4.5M_\\odot/L_\\odot$ for a Kroupa IMF), in agreement with stellar dynamical estimates based on our new velocity dispersions. In the inner arcsecs the luminosity-weighted age drops to 4-8 Gyr, while the metallicity increases to above 3 times the solar value. Starting from 6 arcmin from the center along the major axis, the mean age drops to $\\le 8$ Gyr, with slight supersolar metallicity ($\\approx +0.1$ dex) and $\\alpha-$element overabundance ($\\approx +0.2$ dex), for a mass-to-light ratio $M/L_R\\le 3M_\\odot/L_\\odot$. Diagnostic diagrams based on the [OIII]/H$\\beta$ and [NI]/H$\\beta$ emission line equivalent widths (EWs) ratios indicate that the gas is ionized by shocks outside 10 arcsec, but an AGN-like ionizing source could be present near the center. We speculate that a gas-rich minor merger happened some 100 Myr ago, causing the observed minor axis gas counterrotation, the recent star formation event, and possibly some nuclear activity.} ", "introduction": "\\label{sec_intro} This is the first of two papers presenting new optical spectra for the bulge of M31 to study its stellar populations and assess its triaxiality through dynamical modeling. Here we present the new data and constrain the stellar populations. In the past 50 years papers studying the dynamics of our neighbour galaxy M31 have been published on a regular basis, discussing gas kinematics, both by optical spectroscopy (Boulesteix et al. \\cite{Boulesteix87}, Pellet \\cite{Pellet76}), and in HI (Kent \\cite{Kent89a}, Braun \\cite{Braun91}, Chemin et al. \\cite{Chemin09} and references therein), stellar kinematics concentrating on the central regions to probe the black hole dynamics (Bender et al. \\cite{Bender05}) or considering the whole bulge (McElroy \\cite{McElroy83}). The data are used to construct dynamical models of the galaxy (Widrow et al. \\cite{Widrow03}, Klypin, Zhao and Somerville \\cite{Klypin02}) and possibly probe the tridimensional distribution of its stellar components. The question of the triaxiality of M31 bulge has been posed early on (Stark \\cite{Stark77}, Gerhard \\cite{Gerhard86}) and is of significant importance for the understanding of M31, but a definitive quantitative modeling of both photometry and kinematics is still missing. A bar could also be present (Athanassoula \\& Beaton \\cite{Athanassoula06}, Beaton et al. \\cite{Beaton07}). Moreover, investigations of the stellar populations of the central regions of M31 through the measurement of Lick indices have been performed (Davidge \\cite{Davidge97}). They indicate the presence of a young and metal rich population in the inner arcsecs of the galaxy. Parallely, studies of the resolved stellar population of the bulge of M31 have assessed that the global stellar population of the M31 bulge must be as old as the bulge of Milky Way, resolving previous claims of younger ages as due to crowding problems (Stephens et al. \\cite{Stephens03}). Two considerations convinced us of the necessity to collect new optical spectroscopic information for the bulge of M31, in addition to the old age of the dataset of McElroy (\\cite{McElroy83}). The first one is the start of PAndromeda, an extensive monitoring campaign of M31 with the PanSTARRS-1 telescope and camera system (Kaiser \\cite{Kaiser04}), that is in principle able to deliver hundreds of pixel lensing events, probing both bulge and disk regions. Detailed stellar population and dynamical models, based on accurate spectral information, are needed to interpret these events as due to a compact baryonic dark matter component (the so-called MACHOs) rather than self-lensing of stellar populations (Kerins et al. \\cite{Kerins01}, Riffeser et al. \\cite{Riffeser06}). The second is the development of new modeling techniques of both simple stellar populations and stellar dynamical systems. On the one hand, the interpretation of Lick indices (Worthey et al \\cite{Worthey94}) in terms of the most recent simple stellar population models (Maraston \\cite{Maraston98}, \\cite{Maraston05}) that take into account the variation of $[\\alpha/$Fe] (Thomas, Maraston and Bender \\cite{TMB03}), allows the accurate determination of the stellar population ages, metallicities and overabundances, and therefore the prediction of stellar mass-to-light ratios. On the other hand, new dynamical modeling codes, like N-MAGIC (De Lorenzi et al. \\cite{DeLorenzi07}) allow the flexible dynamical modeling of triaxial structures, optimally exploiting the information contained in the line-of-sight velocity distributions that modern programs for the analysis of the galaxy optical spectra are able to extract (Bender, Saglia and Gerhard \\cite{BSG94}), well beyond the mean velocities and velocity dispersions of McElroy (\\cite{McElroy83}). In the following we discuss our new spectroscopic observations of the bulge of M31. A future paper (Morganti et al. \\cite{Morganti09}) will report on the dynamical modeling. In Sect. \\ref{sec_obs} we present the observations and the data reduction. In Sec. \\ref{sec_kinematics} we derive the stellar and gas kinematics and the strengths of the absorption and emission lines. In Sec. \\ref{sec_mod} we discuss the modeling of the new spectroscopic data. We analyze the stellar population in Sect. \\ref{sec_stelpop} and discuss previous axisymmetric dynamical models of the bulge of M31 in Sect. \\ref{sec_dyn}. Sect. \\ref{sec_ionize} considers the possible excitation sources compatible with the observed emission line EW ratios. We draw our conclusions in Sect. \\ref{sec_conc}. ", "conclusions": "\\label{sec_conc} We have presented new optical spectroscopic observations of the bulge of M31. We have measured the stellar and gas kinematics, emission line strength ratios and absorption Lick indices profiles along 6 position angles out to distances of 5 arcmin from the center. Along the major axis we probed regions out to 8 arcmin. We show that the old kinematics of McElroy (\\cite{McElroy83}) provides too small velocity dispersions (by up to 30\\%), therefore biasing the dynamical modeling towards too small bulge stellar masses (by a factor 2). Moreover, the new higher averaged velocity dispersion predicts a mass for the central supermassive black hole of M31 that is only a factor 2 below what measured. The new velocity dispersion profiles are now in better agreement with axisymmetric dynamical models with large bulge mass-to-light ratio (Widrow et al. \\cite{Widrow03}), that now match the values derived from stellar population models ($M/L_R\\approx 4-4.5 M_\\odot/L_\\odot$, see below). This implies an upward revision of the predicted self-lensing microlensing event rate of Widrow et al. (\\cite{Widrow03}), and Riffeser et al. (\\cite{Riffeser06}), that are based on lower stellar mass-to-light ratios. The inner ($r\\le 100$ arcsec) bulge is slowly rotating, with a $V/\\sigma\\approx 0.2$. At distances from the center larger than $\\approx 100$ arcsec the measured kinematics becomes increasingly influenced by the rapidly rotating stellar disk. Therefore, the observed variation of the kinematic position angle is suggestive of bulge triaxiality, but needs a proper dynamical modeling of both disk and bulge components to be quantified. The measured gas kinematics confirms the well studied large scale disk rotation. However, a more complex structure, with gas minor axis counter-rotation, is detected in the inner bulge. This might be evidence for a (recent) minor merger, possibly connected to the younger stellar population detected in the inner arcsecs of the galaxy discussed below. The analysis of eight Lick index profiles shows that the bulge of M31 is old, of solar-metallicity and a factor 2 overabundant in $\\alpha$-elements, in agreement with studies of its resolved stellar component (Stephens et al. \\cite{Stephens03}). The line indices and stellar population parameters appear approximately constant on circles, i.e. their isocontours are rounder than the galaxy isophotes, as seen in many ellipticals and bulges (Kuntschner et al. \\cite{Kuntschner06}; Falc\\'on-Barroso et al. \\cite{Falcon06}). Together with the derived old ages, this confirms that the stellar disk out to 5 arcmin from the center is either old (similar to what found for other spiral galaxies, Peletier and Balcells \\cite{Peletier96}) or not sufficiently probed by the spectral features considered here. However, we do detect smaller ages ($\\le 8$ Gyr) along the major axis at distances $\\ge 6$ arcmin. The u-g SLOAN colors predicted from our stellar population analysis match reasonably well with the observed ones. The redder colors g-r and r-i are systematically offset, a well known problem of the flux calibration of current simple stellar population models (Maraston et al. \\cite{Maraston09}). The derived mass-to-light ratios (in the Johnson R band and with a Kroupa IMF we get $M/L\\approx 4-4.5$) agree with the dynamical estimates (see above). They drop to $\\le 3 M_\\odot/L_\\odot$ along the major axis at distances $\\ge 6$ arcmin, where the disk light starts to dominate. In the inner arcsecs the situation changes and a population with a light-weighted younger age ($\\approx 8$ Gyr inside the seeing disk or 2 arcsec, with values as low as 4 Gyr) and metal richer ($\\approx 3$ times solar) appears. This agrees with the findings of Davidge (\\cite{Davidge97}). In addition, the emission line EW ratios [OIII]/H$\\beta$ increase in this region. Their values are compatible with being excited by shocks in the main body of the bulge, but near the center they increase to levels pointing to the presence of an AGN-like photo-ionizing source. Combined with the detection of counterrotating gas along the minor axis of the galaxy (see above), this suggests that a gas-rich minor merger probably happened some 100 Myr ago, that triggered an episode of star formation and possibly boosted the nuclear activity of the central supermassive black hole of M31. We estimate how broad a range of star burst ages and masses can be to give the measured mean value of 8 Gyr in the inner 2 arcsec, when superimposed to the old bulge stars background. To this purpose we compute composite spectra of an old (12.6 Gyr) plus a young (from 100 Myr to 4 Gyr) simple stellar population, using the Vazdekis (\\cite{Vazdekis99}) library, and measure their SSP ages through the analysis of the Lick indices observed here. We find that global ages smaller than 8 Gyr are found when considering a young component younger than $\\approx 600$ Myr and a mass fraction lower than 10\\%. Higher mass fractions are possible for older ages. From Kormendy and Bender (\\cite{Kormendy99}), inside an aperture of 2 arcsec diameter we measure a V mag of 12.5, or $5\\times 10^6 L_\\odot$. In this region we estimate $M/L_V\\approx 4 M_\\odot/L_\\odot$ and therefore an enclosed mass of $2\\times 10^7 M_\\odot$. As a consequence, $\\approx 10^6 M_\\odot$ of some 100 Myr old stars would be needed. Note that in the inner nucleus of M31, at fractions of an arcsec, a disk of 200 Myr old stars is found (Bender et al. \\cite{Bender05}), with a mass of $\\approx 4200 M_\\odot$. Of course the spatial resolution of our spectra is too low to probe this scale. Moreover, our result, based on the H$\\beta$ line as an age tracer, depends heavily on the details of the emission correction and might be affected by the bad column's interpolation (see Sect. \\ref{sec_obs}). Spectra of the higher order Balmer lines, taken with better seeing, are needed to improve our conclusions. In a second paper (Morganti et al. \\cite{Morganti09}), a dynamical model of the data presented here, that takes into account the contributions of the bulge and disk components, will assess in a quantitative way the bulge triaxiality issue and will give a proper estimate of the microlensing event rates due to self-lensing." }, "0910/0910.2715_arXiv.txt": { "abstract": "We present the first results of a program to characterize the disk and envelope structure of typical Class~0 protostars in nearby low-mass star forming regions. We use \\textit{Spitzer} IRS mid-infrared spectra, high resolution CARMA 230~GHz continuum imaging, and 2-D radiative transfer models to constrain the envelope structure, as well as the size and mass of the circum-protostellar disk in Serpens FIRS~1. The primary envelope parameters (centrifugal radius, outer radius, outflow opening angle, and inclination) are well constrained by the spectral energy distribution (SED), including \\textit{Spitzer} IRAC and MIPS photometry, IRS spectra, and 1.1~mm Bolocam photometry. These together with the excellent $uv$-coverage ($4.5-500$~k$\\lambda$) of multiple antenna configurations with CARMA allow for a robust separation of the envelope and a resolved disk. The SED of Serpens FIRS~1 is best fit by an envelope with the density profile of a rotating, collapsing spheroid with an inner (centrifugal) radius of approximately 600~AU, and the millimeter data by a large resolved disk with $\\mdisk\\sim1.0~ \\msun$ and $\\rdisk\\sim 300$~AU. These results suggest that large, massive disks can be present early in the main accretion phase. Results for the larger, unbiased sample of Class~0 sources in the Perseus, Serpens, and Ophiuchus molecular clouds are needed to determine if relatively massive disks are typical in the Class~0 stage. ", "introduction": "Protostars build up their mass by accreting material from a dense protostellar envelope, presumably via a rotationally supported circum-protostellar accretion disk. Disk formation is a natural result of collapse in a rotating core, but it is not known how soon after protostellar formation the disk appears, or how massive it is at early times. Theory suggests that centrifugally supported disks should start out small (radius less than $10$~AU), and thus low mass, and grow with time \\citep*{tsc84}. Unstable or magnetically supported disks, however, could be much larger (radii up to 1000~AU in the magnetically supported case; \\citealt{gs93}), and thus more massive at early times. The remnants of these protostellar accretion disks are easily observed in more evolved phases (e.g. TTauri stars), but given that the majority of mass is accreted during earlier embedded phases, understanding disks at early times is critical. Directly observing disks during this main accretion phase is quite difficult, however, as they are hidden within dense, extincting protostellar envelopes. The structure of the envelope at small radii is another important characteristic of main accretion phase protostars that is similarly difficult to directly observe. Disk growth or the presence of a binary companion may clear out the inner region of the envelope early on, as inferred for the binary Class~0 source IRAS 16293-2422 by \\citep{jorg05}. There has been a recent push toward detecting disks in more embedded objects, with many now known and roughly characterized in Class I protostars \\citep[e.g.][]{loon00,jorg05b,eisner05,aw07}, and a few detected in the earlier Class 0 stage \\citep[e.g.][]{chan95,harv03,brown00,loon00}. \\footnote{We use definitions of Class 0, Class I, and Class II \\citep{andre94} based on the bolometric temperature \\citep{ml93,chen95}: $\\tbol \\le 70$~K (Class~0); 70 K$< \\tbol \\le 650$~K (Class~I); 650 K$<\\tbol \\le 2800$ K (Class~II). We further assume that classes correspond to an evolutionary sequence \\citep[e.g.][]{rob06}: in Class~0 the protostar has accreted less than half its final mass ($M_{*} < \\menv$), in Class~I $M_* > \\menv$, and in Class~II the envelope has dispersed, leaving only a circum-stellar disk.} Most previous detailed studies have been limited to the most well-known or brightest Class 0 sources, however, due to instrumental limitations and a lack of large unbiased target samples. The ongoing Submillimeter Array survey of low-mass protostars \\citep{jorg07,jorg09} is a notable exception. With recent large surveys of nearby molecular clouds at mid-infrared and (sub)millimeter wavelengths it is now possible to define complete samples of Class 0 protostars based on luminosity or envelope mass limits \\citep[e.g.][]{hatch07,jorg08,dun08,enoch09,evans09}. \\begin{figure*}[!ht] \\vspace{-0.2in} \\begin{center} \\includegraphics[width=4.8in]{f1.eps} \\vspace{-0.2in} \\caption {\\textit{Spitzer} $24~\\micron$ image of the immediate environment of Serpens FIRS~1, in the Serpens main core. Bolocam 1.1~mm continuum contours are overlaid. Submillimeter source designations for the brightest \\citet{casali93} sources are indicated. The nearest embedded protostar to FIRS~1 is approximately $45\\arcsec$ or $11000$~AU away (Ser-emb 12; \\citealt{enoch09}), and the nearest YSO is $25\\arcsec$ or 6000~AU away (c2d 142; \\citealt{harv07}). \\label{genfig}} \\end{center} \\end{figure*} We have recently begun a campaign to characterize disk properties in a large, uniform sample of Class 0 protostars in nearby low mass star-forming regions (M. Enoch et al. 2009, in preparation). Our study is based on the complete (to envelope masses $\\gtrsim0.1~\\msun$) sample of 39 Class 0 protostars in the Serpens, Perseus, and Ophiuchus molecular clouds, identified by \\citet{enoch09} by comparing large-scale \\textit{Spitzer} IRAC and MIPS and Bolocam 1.1~mm continuum surveys of the three clouds. Combining Spitzer IRS mid-infrared (MIR) spectra and high resolution CARMA 230~GHz continuum imaging with radiative transfer modeling of this sample will help to address several fundamental questions about the structure and evolution of the youngest protostars: 1) How soon after the initial collapse of the parent core does a circum-protostellar disk form? 2) What fraction of the total circum-protostellar mass resides in the disk, and does this fraction vary with time? 3) Are large ``holes'' in the inner envelope, such as that found for IRAS~16293 by \\citet{jorg05}, common at early times? MIR spectra and millimeter maps provide complementary approaches to these questions. The amount of flux escaping at $\\lambda \\lesssim 50~\\micron$ from deeply embedded sources is very sensitive to the opacity close to the protostar, and thus the envelope structure \\citep[e.g.][]{jorg05}. While the MIR flux is insensitive to disk properties, high resolution millimeter continuum mapping can directly detect emission from dust grains in the disk. Millimeter observations with excellent $uv$-coverage, combined with radiative transfer models, can separate the disk from the envelope and constrain the disk mass and size. Our ultimate goal is to characterize the disk mass, size, and inner envelope structure of typical low-mass Class~0 protostars, and to quantify any trends with evolutionary indicators. In this initial paper we present results for Serpens FIRS~1, a well known Class 0 source, which will serve as a test case for the full program. ", "conclusions": "" }, "0910/0910.0350.txt": { "abstract": "{ An overview of some recent developments in inhomogeneous models is presented. As the volume and precision of cosmological data improves, it will become more and more essential to understand the non-linear behaviour of the Einstein field equations. This requires the study of exact inhomogeneous solutions, including their density distributions, their evolution, their geometry, and their causal structure. Observations are strongly affected by the detailed geometry and evolution of a model, and therefore interpretation of observations depends on understanding them. It is generally assumed the universe is homogeneous if averaged over large enough scales, but to actually prove this is so, will require the assumption to be relaxed, and a rigorous inhomogeneous approach to be applied. Though the \\LT\\ metric has long been used for models of spherical inhomogeneities, there have been a number of new results, including a variety of methods for creating models with specific properties, and their application to cosmic structures on several different scales. Interest in the Szekeres metrics is on the increase, and the quasi-spherical metric was recently used to model specific cosmic structures for the first time. The quasi-planar and quasi-hyperspherical metrics have been hardly studied until recent work invesigated their physical and geometric properties. There is enormous scope for work with these metrics. } \\FullConference{5th International School on Field Theory and Gravitation,\\\\ April 20-24, 2009\\\\ Cuiab\\'{a} city, Brazil} \\tableofcontents \\begin{document} ", "introduction": "Why study inhomogeneous models? The real universe is very lumpy. To properly understand what we see, we should apply all possible methods: perturbation theory, $N$-body newtonian simulations, and exact inhomogeneous metrics --- each has its domain of validity. Inhomogeneous metrics have the advantage that they are fully non-linear and relativistic solutions of the Einstein field equations (EFEs). The assumption of homogeneity has become so well established, that it has become all-pervasive. But now, with so much data coming in, it's time to test homogeneity. Cosmological data reduction relies heavily on the Robertson-Walker (RW) metric --- we need to beware of a circular argument. It will be a significant challenge to check which of our well-known results actually depend on the assumption of homogeneity, and to re-derive them all without that assumption. Here I will present a selection of results in inhomogeneous cosmology, especially work done with Lu, McClure, Krasi\\'{n}ski, Bolejko, C\\'{e}l\\'{e}rier, Alfedeel, Mustapha, Ellis and others, but I won't try to be comprehensive. I'll attempt to provide the basics, and thereby promote the use of inhomogeneous metrics for the study of cosmological problems. Inhomogeneous metrics will become more important as the amount and accuracy of cosmological data increases, and more precise analysis is needed, so there are plenty of opportunities for good research. ", "conclusions": "The universe is of course very inhomogeneous on many scales. To fully understand how these structures evolve, and properly analyse our observations will require the non-linearity of exact inhomogeneous metrics. Up to now, homogeneity has been assumed, and was key to making progess. In the age of precision cosmology, we should thoroughly test this assumption and quantify how good an approximation it is on each scale. Nearly all data analysis assumes the RW metric. To be sure we avoid circular arguments, there is an urgent need to re-do all calculations in a general non-homogeneous metric. The methods of inhomogeneous cosmology will be an essential component of this endeavour. \\LT\\ models have produced a wide variety of interesting results, and the investigations are far from exhausted. The Szekeres models have a lot of flexibility, and can be used to model quite complex structures --- but they have been very little investigated. There are plenty of opportunities for good research. \\appendix" }, "0910/0910.1716_arXiv.txt": { "abstract": "{The galaxy cluster XMMU~J2235.3$-$2557 (hereafter XMM2235), spectroscopically confirmed at $z=1.39$, is one of the most distant X-ray selected galaxy clusters. It has been at the center of a multi-wavelength observing campaign with ground and space facilities.} {We characterize the galaxy populations of passive members, the thermodynamical properties and metal abundance of the hot gas, and the total mass of the system using imaging data with HST/ACS ($i_{775}$ and $z_{850}$ bands) and VLT/ISAAC (J and K$_S$ bands), extensive spectroscopic data obtained with VLT/FORS2, and deep (196\\,ks) \\Chandra observations. } {\\Chandra data allow temperature and metallicity to be measured with good accuracy and the X-ray surface brightness profile to be traced out to 1\\arcmin\\ (or 500 kpc), thus allowing the mass to be reliably estimated. Out of a total sample of 34 spectroscopically confirmed cluster members, we selected 16 passive galaxies (without detectable [OII]) within the central 2\\arcmin\\ (or 1 Mpc) with ACS coverage, and inferred star formation histories for subsamples of galaxies inside and outside the core by modeling their spectro-photometric data with spectral synthesis models.} {\\Chandra data show a regular elongated morphology, closely resembling the distribution of core galaxies, with a significant cool core. We measure a global X-ray temperature of $kT = 8.6_{-1.2}^{+1.3}$ keV (68\\% confidence), which we find to be robust against several systematics involved in the X-ray spectral analysis. By detecting the rest frame 6.7 keV Iron K line in the \\Chandra spectrum, we measure a metallicity $Z= 0.26^{+0.20}_{-0.16} \\, Z_\\odot$. In the likely hypothesis of hydrostatic equilibrium, we obtain a total mass of $M_{\\rm tot}(<1\\,{\\rm Mpc})= (5.9\\pm 1.3)\\times 10^{14}\\,M_\\odot$. By modeling both the composite spectral energy distributions and spectra of the passive galaxies in and outside the core, we find a strong mean age radial gradient. Core galaxies, with stellar masses in excess of $10^{11} M_\\odot$, appear to have formed at an earlier epoch with a relatively short star formation phase ($z=5-6$), whereas passive galaxies outside the core show spectral signatures suggesting a prolonged star formation phase to redshifts as low as $z\\approx 2$. } {Overall, our analysis implies that XMM2235 is the hottest and most massive bona--fide cluster discovered to date at $z>1$, with a baryonic content, both its galaxy population and intracluster gas, in a significantly advanced evolutionary stage at 1/3 of the current age of the Universe. } ", "introduction": "Over the past two decades, considerable effort has been devoted to discovering ever more distant galaxy clusters using different observational methods \\citep[e.g.][for a review]{2002ARA&A..40..539R}. These studies have been traditionally motivated by cosmological applications of the cluster abundance at high redshift \\citep[e.g.][for a review]{2005RvMP...77..207V}, and also by the use of clusters as laboratories to investigate galaxy evolution. Clusters provide a convenient and efficient way of studying large populations of early-type galaxies, which provide stringent tests on galaxy evolution models in the current hierarchical formation paradigm \\citep{2006ARA&A..44..141R}, because they are the most massive galaxies with the oldest stellar populations (at least out to $z\\sim\\! 2$). Clearly, the higher the redshift the stronger the leverage on theoretical models. In addition, galaxy properties in clusters can be contrasted with those in field surveys, which have multiplied in recent years, thus extending the baseline over which environmental effects can be studied. X-ray selection of clusters has been central in these studies, as it naturally provides gravitationally bound systems (as opposed to simple overdensities of galaxies) with a relatively simple selection function. Using ROSAT serendipitous surveys, supported by near IR imaging and spectroscopy with 8-10m class telescopes, the redshift envelope was pushed to $z=1.3$, with only 5 clusters discovered at $z>1$, approximately one per square degree \\citep{2002ARA&A..40..539R}. The extension of the same technique to \\XMM serendipitous surveys has led to the discovery of two clusters at $z>1.3$ to date, XMMU J2235.3$-$2557 at $z=1.39$ \\citep{2005ApJ...623L..85M} (hereafter M05) and XMMXCS~J2215.9$-$1738 at $z=1.46$ \\citep{2006ApJ...646L..13S}. See also \\citet{2008A&A...487L..33L} for a newly discovered, high X-ray luminosity massive cluster at $z\\sim\\! 1$. With the advent of the {\\it Spitzer} observatory, an alternative and efficient way to unveil distant clusters over large areas (as red galaxy overdensities in the IRAC and optical bands) has been developed. This technique has given notable results, with three clusters spectroscopically confirmed at $z>1.3$ in the IRAC Shallow Survey \\citep[ISCS;][]{2008ApJ...684..905E} and one in the SpARCS survey \\citep{2009ApJ...698.1943W}. Detailed investigations of galaxy populations in the X-ray luminous clusters at $z>1$ have been conducted with HST/ACS in combination with the VLT and the Keck telescopes. The study of clusters at $z=1.10,\\, 1.24,\\, 1.26,\\, 1.27$ \\citep{2006ApJ...644..759M, 2007ApJ...663..164D, 2009ApJ...690...42M} and the aforementioned systems at $z=1.39$ \\citep{2008A&A...489..981L} and $z=1.46$ \\citep{2009ApJ...697..436H}, have revealed tight red sequences for the early-type galaxies, with scatters only marginally larger than those in local clusters, implying that most of their stellar mass was assembled at $z>3$ with passive evolution thereafter. While a change in the morphology-density relation of early type galaxies (E+S0 galaxies) has been observed at $z\\sim\\! 1$, elliptical galaxies still dominate the cluster galaxy population up to $z\\sim\\! 1.2$ \\citep{2005ApJ...623..721P,2007ApJ...670..190H}. A comparison of cluster and field early-types of similar stellar mass has revealed a mild but significant difference between the star formation histories in the two environments \\citep[e.g.][]{2008A&A...488..853G,2008arXiv0806.4604R,2007ApJ...655...30V}, a result which is predicted by current hierarchical galaxy formation models \\cite[e.g.][]{2008ApJ...685..863M}. To date, such comparative studies can only be carried out at $z\\lesssim 1.4$, whereas the existence of a substantial population of old, massive, passively evolving early-type galaxies in the field is now well established up to $z\\sim 2$ \\citep[][]{2006ApJ...649L..71K,2008A&A...482...21C}. By pushing cluster studies to higher redshifts, where evolutionary time scales become comparable to the age of the Universe at these redshifts, one would expect to detect significant evolutionary effects. However, as discussed in this paper, this has not been the case so far, even after probing two-thirds of the look-back time, not only for the galaxy populations but also for the thermodynamical properties and chemical enrichment of the hot gas measured with follow-up \\Chandra observations in $z>1$ clusters. Our current understanding is that relations, such as the red sequence in the color-magnitude diagram, the morphology-density relation, the $L_X-T_X$ relation for the intracluster gas, emerge at $z\\lesssim 2$. For example, the study of a proto-cluster at $z=2.16$ identified around a powerful radio galaxy provided evidence of a forming red sequence \\citep{2008ApJ...680..224Z,2007MNRAS.377.1717K}, which likely takes 1-2 Gyrs to form \\citep{2008A&A...488..853G}. It is unfortunate that such a transition in the assembly process of galaxy clusters seems to occur in a redshift range where spectroscopic observation are particularly difficult. In this spirit, we have carried out a multi-wavelength study of one the most distant clusters known, XMMU J2235.3$-$2557 (hereafter XMM2235) at $z=1,39$, which was the first distant cluster confirmed (M05) as part of the on-going \\XMM Distant Cluster Project \\citep[XDCP,][]{2005Msngr.120...33B,2008arXiv0806.0861F}. In this paper, we use spectro-photometric observations of XMM2235 in the optical/near-IR and X-ray bands to characterize its galaxy population (particularly the passive spectroscopic members) and the thermodynamic status of the hot intracluster gas and to measure its total mass. In section \\ref{sec:VLTHST}, we present the VLT and HST data used in this paper, including an extensive spectroscopic campaign which yielded 34 confirmed cluster members. In section \\ref{sec:SFH}, we model the underlying stellar populations of passive galaxies and constrain their star formation histories. In section \\ref{sec:Xray}, we present deep \\Chandra observations of XMM2235, the methods of analysis and the resulting measurements of its temperature, metallicity and mass. In section \\ref{sec:results}, we discuss the results. $H_0=70\\ {\\rm km}\\ {\\rm s}^{-1}\\,{\\rm Mpc}^{-1}, \\,\\Omega_m=0.3, \\, \\Omega_\\Lambda=0.7$ are adopted throughout this paper. In this cosmology, 1\\arcmin\\ on the sky corresponds to 0.5 Mpc at $z=1.39$ ", "conclusions": "\\label{sec:results} We presented a combination of spectro-photometric data from VLT and HST, as well as deep \\Chandra observations of the X-ray selected cluster XMMU J2235.3$-$2557 at $z=1.39$ and used them to characterize the galaxy populations of passive members, the thermodynamical properties of the hot gas and the total mass of the system. Despite surface brightness dimming, due to the high redshift of the cluster, the \\Chandra data show extended X-ray emission out to $\\sim\\! 0.5$ Mpc. The X-ray emission has a regular morphology and is clearly elongated in the same way as the distribution of red passive galaxies in the core as well as the major axis of the BCG. An excess of emission in the inner 50 kpc is clearly detected and naturally interpreted as a cool core. The spectral analysis of the \\Chandra data reveals that XMM2235 has a global temperature of $kT = 8.6_{-1.2}^{+1.3}$ keV (68\\% confidence), which we find robust against several systematics involved in the X-ray spectral analysis. If we assume hydrostatic equilibrium, a well justified condition given the relaxed appearance and the canonical $L_X/T_X$ ratio of the cluster, and an isothermal gas distribution, the X-ray surface brightness profile yields a total mass at large radii ($r\\gg 100$ kpc) of $M_{\\rm tot}(1$. These findings are corroborated by the weak lensing analysis from deep HST/ACS data \\citep{2009ApJ...704..672J} which provides a mass profile in very good agreement with the X-ray measurement. The presence of a significant cool core is additional evidence of the advanced dynamical state of the cluster \\citep[e.g.][]{1994ARA&A..32..277F}. In characterizing the passive galaxy population of XMM2235, we have extended the analysis of the ACS Intermediate Redshift Cluster Survey at $0.8 5$) also suggest differences between the dust extinction properties in higher redshift galaxies and the local universe \\citep[e.g.][]{mso+04,td01}. However, the degeneracy that exists between the best-fit GRB spectral index and the host galaxy's total-to-selective extinction means that to accurately determine the host galaxy's extinction law properties, good quality, broadband data are needed, preferentially stretching out into the negligibly extinguished far infrared (FIR) wavelength bands. In the current era of \\swift\\ and rapid-response ground-based telescopes, prompt arcsecond GRB positions have provided a wealth of high quality, early-time X-ray, ultraviolet (UV), optical and NIR data. Accurate soft X-ray absorption measurements are now available for the large fraction of GRBs \\citep{crc+06,bk07,gnv+07,smp+07,ebp+09}, and well-sampled, high signal-to noise SEDs are providing strong constraints on the best-fit extinction law models \\citep[e.g.][]{pbb+08}. There have now been some examples of GRB host galaxies with the 2175~\\AA\\ absorption feature (e.g. GRB~070802, El{\\'{\\i}}asd{\\'o}ttir et al. 2009, Kr{\\\"u}hler et al. 2008; GRB~080607, Prochaska et al. 2009), as well as GRB host galaxies with \\Rv\\ values larger than the mean SMC, LMC and MW values \\citep[e.g.][]{pbb+08}, a possible indicator of grey dust, as suggested to be present in some GRB host galaxies \\citep[e.g.][]{sff03}. However, such analysis on the detailed properties of GRB extinction curves are typically still only possible for a handful of well-sampled, bright GRBs (e.g. GRB050525A, Heng et al. 2008; GRB~061126, Perley et al. 2009; GRB~070802, El{\\'{\\i}}asd{\\'o}ttir et al. 2009, Kr{\\\"u}hler et al. 2008; GRB~080607, Prochaska et al. 2009). In \\citet{smp+07} we used X-ray and UV/optical simultaneous observations taken with the X-ray Telescope \\citep[XRT;][]{bhn+05} and Ultraviolet and Optical Telescope \\citep[UVOT;][]{rkm+05} onboard \\swift\\ \\citep{gcg+04} to analyse the SEDs for a sample of 7 GRBs. The dust extinction in the GRB host galaxies was modelled on the mean SMC, the LMC and the Milky Way extinction curves using the parameterisations given in \\citet{pei92}, which cover a range in 2175~\\AA\\ bump strengths and \\Rv\\ values. The SMC and LMC extinction curves were found to provide the best-fit model for the majority of the sample, in agreement with previous studies \\citep[e.g.][]{sfa+04,kkz06,sww+07}. However, we also found that, although the gas-to-dust ratio in \\swift\\ GRB host galaxies were typically larger than those of the Milky Way and Magellanic Clouds, the weighted mean was within 90\\% confidence of the Magellanic Clouds and Milky Way X-ray absorption to optical extinction ratios. In this paper we aim to further our previous work, using a larger sample of 28 GRBs, and increasing the wavelength range of the afterglow SEDs to better constrain the absorption and extinction within the GRB host galaxy. In \\citet{smp+07} \\swift\\ data alone were used to produce the SEDs, and UVOT data with a rest-frame wavelength $\\lambda < 1215$~\\AA\\ were not included in the SED fits in order to avoid the absorption caused by the Lyman forest being confused for dust-extinction. In this paper we now model the absorption resulting from the Lyman forest such that all rest frame UV data redward of the Lyman edge is included in our spectral analysis. Furthermore, we also include additional ground-based NIR data if available, further increasing the spectral range of the SEDs and the degrees of freedom of the spectral fits. This provides better sampled SEDs and extends the redshift range within our sample, which was previously restricted to $z<1.7$ to ensure that the SED modelling was sufficiently well constrained within the optical wavelength range. In $\\S$\\ref{sec:analysis} we present the new, extended GRB sample and describe the X-ray, UV/optical and NIR data reduction and analysis, and in $\\S$\\ref{sec:model} we describe the models used to fit the data. We present the results of our spectral modelling in $\\S$\\ref{sec:results} followed by an analysis of the possible selection effects and systematic biases that may be present in our work in $\\S$\\ref{sec:seleffs}. A discussion on the implications of our findings is presented in $\\S$\\ref{sec:disc}, and our conclusions are summarised in $\\S$\\ref{sec:cons}. Throughout the paper temporal and spectral indices, $\\alpha$ and $\\beta$, respectively, are denoted such that $F(\\nu,t)\\propto \\nu^{-\\beta}t^{-\\alpha}$, and all errors are $1\\sigma$ unless specified otherwise. %====================================TABLE 1==================================== %=============================================================================== \\begin{table*} \\caption{Table listing the 28 GRBs in our sample with their redshift, Galactic column density and visual extinction in the line-of-sight to the GRB, the corresponding SED epoch, the UV, optical, and NIR band passes included in the GRB afterglow SED, and the rest frame coverage of the SED.\\label{tab:sample}} \\begin{tabular}{@{}lllllll} \\hline GRB & z & $N_{H,X}$(Gal) & $A_V$(Gal) & Epoch & UV/optical/NIR bandpasses & Restframe Band \\\\ & & ($10^{21}$~\\invsqrcm) & & (s) & & Coverage (\\AA) \\\\ \\hline\\hline 050318 & 1.44$^a$ & 0.28 & 0.05 & T$^\\dag$+3600 & v,b,u & 1260--2400 \\\\ 050319 & 3.24$^b$ & 0.11 & 0.03 & T+20,000 & $I^{1,2},R^{1-3}$,v,b & 920--2090 \\\\ 050525A & 0.606$^c$ & 0.91 & 0.29 & T+20,000 & $K^{4,5},H^4,J^{4,6},I^{4,8},R^{4,9,10}$,v,b,u,w1,m2,w2 & 1000-14540 \\\\ 050730 & 3.968$^d$ & 0.30 & 0.16 & T+10,000 & $K^{11},J^{11},I^{11},R^{11}$,v,b & 780--4700 \\\\ 050802 & 1.71$^e$ & 0.18 & 0.06 & T+20,000 & $I^{12},R^{12}$,v,b,u,w1,m2,w2 & 590--3270 \\\\ 050820A & 2.6147$^f$ & 0.47 & 0.14& T+10,000 & $J^{13},z^{14},I^{14},R^{14},g^{14}$,v,b,u,w1 & 620--3710 \\\\ 050922C & 2.198$^g$ & 0.54 & 0.32 & T+20,000 & $R^{15-20}$,v,b,u,w1,m2 & 620--2360 \\\\ 051109A & 2.346$^h$ & 1.61 & 0.59 & T+5000 & $K^{21},H^{21},J^{21},I^{22},R^{23,24}$,v,b,u,w1 & 670--6980 \\\\ 060124 & 2.296$^i$ & 0.92 & 0.42 & T+100,000 & $I^{25},R^{25}$,v,b & 1180--2690 \\\\ 060206 & 4.048$^j$ & 0.09 & 0.04 & T+10,000 & $K^{26},H^{26},J^{26},R^{27-29}$,v,b & 770--4630\\\\ 060418 & 1.49$^k$ & 0.92 & 0.69 & T+5000 & $K^{30},H^{30},J^{30},z^{30,31},I^{32},R^{33}$,v,b,u,w1,m2 & 800--9380 \\\\ 060502A & 1.51$^l$ & 0.30 & 0.10 & T+5000 & $R^{34}$,v,b,u,w1 & 900--3010 \\\\ 060512 & 0.4428$^m$ & 0.14 & 0.05 & T+10,000 & $Ks^{35},J^{36},R^{37,38}$,v,b,u & 2130--16180 \\\\ 060526 & 3.21$^n$ & 0.55 & 0.21 & T+20,000 & $J^{39},I^{39,40},R^{40-47}$,v,b & 930--3180 \\\\ 060605 & 3.711$^o$ & 0.51 & 0.15 & T+10,000 & $R^{48-53}$,v,b & 830--1600 \\\\ 060607A & 3.082$^p$ & 0.27 & 0.09 & T+10,000 & $H^{30},i^{54},r^{54},g^{54}$,v,b,u & 750--4200 \\\\ 060714 & 2.71$^q$ & 0.61 & 0.24 & T+5000 & $R^{55}$,v,b & 1050--2040 \\\\ 060729 & 0.54$^r$ & 0.48 & 0.17 & T+70,000 & $R^{56}$,v,b,u,w1,m2,w2 & 1040--4900 \\\\ 060904B & 0.703$^s$ & 1.21 & 0.53 & T+5000 & $K^{57},J^{57},I^{57 },R^{58,59}$,v,b,u,w1,m2,w2 & 940--13710 \\\\ 060908 & 2.43$^t$ & 0.27 & 0.09 & T+5000 & $R^{60,61}$,v,b,u,w1 & 660--2200 \\\\ 060912 & 0.937$^u$ & 0.42 & 0.16& T+1500 & v,b,u,w1,m2 & 1030--3020 \\\\ 061007 & 1.262$^v$ & 0.21 & 0.06 & T+600 & $i^{62},R^{62}$,v,b,u,w1,m2,w2 & 710--3850 \\\\ 061121 & 1.314$^w$ & 0.51 & 0.14 & T+10,000 & $I^{62-64},R^{65}$,v,b,u,w1,m2,w2 & 690--3830 \\\\ 061126 & 1.159$^x$ & 1.00 & 0.56 & T+2000 & $K^{66},J^{66},I^{66,67},R^{66,68,69}$,v,b,u,w1,m2 & 920--10820 \\\\ 070110 & 2.352$^y$ & 0.18 & 0.04 & T+10,000 & $R^{70}$,v,b,u & 920--2250 \\\\ 070318 & 0.836$^z$ & 0.25 & 0.05 & T+1500 & v,b,u,w1,m2,w2 & 870--3190 \\\\ 070411 & 2.954$^\\ddag$ & 2.63 & 0.88 & T+500 & $R^{71,72}$,v,b & 990--1900 \\\\ 070529 & 2.4996$^\\S$ & 1.90 & 0.93 & T+600 & v,b,u,w1 & 640--1670 \\\\ \\hline \\end{tabular} \\begin{list}{}{} \\item[] $^a$ \\citet{bm05}; $^b$ \\citet{fhj+05}; $^c$ \\citet{fcb+05}; $^d$ \\citet{sve+05}; $^e$ \\citet{fsj+05}; $^f$ \\citet{lve+05}; $^g$ \\citet{jfp+05a}; $^h$ \\citet{qhr+05}; $^i$ \\citet{pft+06}; $^j$ \\citet{pwf+06}; $^k$ \\citet{dfp+06} $^l$ \\citet{cpf+06}; $^m$ \\citet{bfk+06}; $^n$ \\citet{bg06}; $^o$ \\citet{fkk+06}; $^p$ \\citet{lvs+06}; $^q$ \\citet{jvf+06d}; $^r$ \\citet{tlj+06}; $^s$ \\citet{fdm+06}; $^t$ \\citet{rjt+06}; $^u$ \\citet{jlc+06c}; $^v$ \\citet{jft+06b}; $^w$ \\citet{bpc06}; $^x$ \\citet{pbb+08}; $^y$ \\citet{jmf+07}; $^z$ \\citet{jfa+07}; $^\\ddag$ \\citet{jmt+07}; $^\\S$ \\citet{bfc07} \\item[]$^\\dag$ T is time at which the BAT triggered on the GRB \\item[] $^1$ \\citet{huk+07}; $^2$ \\citet{krm07}; $^3$ \\citet{wvw+05}; $^4$ \\citet{cb05}; $^5$ \\citet{rg05}; $^6$ \\citet{fhs+05}; $^7$ \\citet{ffc+05}; $^8$ \\citet{ytk05}; $^9$ \\citet{hhg+05}; $^{10}$ \\citet{mbs05a}, $^{11}$ \\citet{pcm+06}; $^{12}$ \\citet{pes+05}; $^{13}$ \\citet{meb+05}; $^{14}$ \\citet{ckh+06}; $^{15}$ \\citet{dp05}; $^{16}$ \\citet{jpt+05b}; $^{17}$ \\citet{ap05}; $^{18}$ \\citet{hkh+05}; $^{19}$ \\citet{pmm+05}; $^{20}$ \\citet{dpf+05}; $^{21}$ \\citet{bbs+05}; $^{22}$ \\citet{tor05}; $^{23}$ \\citet{mwp+05}; $^{24}$ \\citet{juc+05}; $^{25}$ \\citet{mbs+07}; $^{26}$ \\citet{apb06}; $^{27}$ \\citet{cvw+07}; $^{28}$ \\citet{sdp+07}; $^{29}$ \\citet{wvw+06}; $^{30}$ \\citet{mvm+07}; $^{31}$ \\citet{nif+06}; $^{32}$ \\citet{cob06a}; $^{33}$ \\citet{kop06}; $^{34}$ \\citet{cof06}, $^{35}$ \\citet{hlm+06}; $^{36}$ \\citet{sdp+06}; $^{37}$ \\citet{cen06a}; $^{38}$ \\citet{mil06}; $^{39}$ \\citet{cob06b}; $^{40}$ \\citet{tgb+06}; $^{41}$ \\citet{bgv+06}; $^{42}$ \\citet{cig+06}, $^{43}$ \\citet{dhm+07}; $^{44}$ \\citet{gtb+06}; $^{45}$ \\citet{kbs+06a}; $^{46}$ \\citet{md06}; $^{47}$ \\citet{rpi+06} $^{48}$ \\citet{kg06}; $^{49}$ \\citet{ksa+06c}; $^{50}$ \\citet{ksa+06b}; $^{51}$ \\citet{mfm+06}; $^{52}$ \\citet{sap06}; $^{53}$ \\citet{zqw+06}; $^{54}$ \\citet{nrc+09}; $^{55}$ \\citet{api06}; $^{56}$ \\citet{qr06}; $^{57}$ \\citet{cb06}; $^{58}$ \\citet{pks+06}; $^{59}$ \\citet{skv06}; $^{60}$ \\citet{act+06}; $^{61}$ \\citet{wtr06}; $^{62}$ \\citet{mmg+07}; $^{62}$ \\citet{cen06b}; $^{63}$ \\citet{cob06c}; $^{64}$ \\citet{tor06a}; $^{65}$ \\citet{uau06}; $^{66}$ \\citet{pbb+08}; $^{67}$ \\citet{tor06b}; $^{68}$ \\citet{smg+06}; $^{69}$ \\citet{wm06}; $^{70}$ \\citet{mjv07}; $^{71}$ \\citet{msd07}; $^{72}$\\citet{klk+07} \\end{list} \\end{table*} %=============================================================================== ", "conclusions": "\\label{sec:cons} In this paper we have presented the results from the spectral analysis of 28 GRB SEDs. We measured the equivalent neutral hydrogen column density and visual extinction at the host galaxy, and found 79\\% of the GRBs in our sample to have a detectable soft X-ray absorption system in the host galaxy, and 71\\% to have a detectable visual dust-extinction system. Using the measured \\nhx/\\av\\ ratios as an indicator of the host galaxy gas-to-dust ratio, we find that GRB host galaxies have gas-to-dust ratios that are typically larger than those measured in the Milky Way and Magellanic Clouds by up to two orders of magnitude. We have investigated several possibilities that could account for the relatively large gas-to-dust ratios in GRB host galaxies. There is no evidence to suggest that the large host galaxy \\nhx/\\av\\ ratios measured in our GRB sample is the result of any systematic error in the way that we measure \\av. One possibility is that dust destruction by the GRB has reduced the visual extinction, \\av, relative to the equivalent neutral hydrogen column density, \\nhx. However, there are currently no observations that clearly show the early time colour evolution expected from dust destruction. Although such observations are limited by the quality and promptness of the data, we also found that the majority of our sample had host dust properties best-fit by the UV steep, SMC extinction law, indicating an abundance of small dust grains in the GRB surrounding environment. In the event of a significant phase of dust destruction, a grey extinction law should be observed, where the differential change in extinction from UV to NIR energy bands is small. The dust probed by our \\av\\ measurements must, therefore, lie in regions of the GRB host galaxy that have not been subjected to significant amounts of dust destruction. For a subset of eight GRBs we were also able to study how the neutral hydrogen column density, \\nh, compared with \\av, and we found \\nh/\\av\\ to extend to both larger and smaller values than those of the Magellanic Clouds and the Milky Way by up to an order of magnitude. The distribution in \\nh/\\av\\ can be accounted for by the competing effects that alter the values of \\nh\\ and \\av. Firstly, differences in the host galaxy metallicities and in the amount of dust destroyed by the GRB will affect the value of \\av. On the other hand, the value of \\nh\\ will be dependent on the amount of photo-ionised hydrogen along the line-of-sight to the GRB. The mean logarithmic metallicity of the GRB sample with both \\nh\\ and \\av\\ measurements is almost 1.0~dex smaller than that of the SMC (0.04~$Z_{\\odot}$), and we would therefore expect the GRB host galaxy \\nh/\\av\\ ratio to be significantly smaller than the SMC \\nh/\\av\\ ratio. The roughly even number of GRBs with smaller and larger \\nh/\\av\\ ratios than the Magellanic Clouds and Milky Way therefore implies that the level of photo-ionised hydrogen along the line-of-sight to the GRB is greater than the fraction of dust destroyed by the GRB. This would suggest that measurements of \\nhx\\ and \\av\\ probe regions of dust and gas much closer to the GRB than \\nh. It has been suggested that differences in the gas-to-dust ratios in galaxies of different types are correlated with the metallicity of the galaxy \\citep[e.g.][]{ddb+07}, whereby smaller metallicity systems have larger gas-to-dust ratios. From a subsample of four GRBs with measured metallicity and a soft X-ray absorption and visual extinction system detected with $90$\\% confidence, together with the Small and Large Magellanic Clouds and Milky Way, we found a strong negative correlation between the \\nhx/\\av\\ ratio and the metallicity, [M/H]. The spearman rank coefficient was -0.89 with $90$\\% confidence. The large \\nhx/\\av\\ ratios measured in GRB host galaxies could, therefore, be an indication of their very low, although broad, range of metallicities. A greater sample of GRB hosts with measured metallicities are needed to verify such a correlation, which if confirmed would suggest that low-metallicity environments are less efficient at forming dust from their metals than high-metallicity galaxies." }, "0910/0910.3655_arXiv.txt": { "abstract": "Observations of transient phenomena in the Universe reveal a spectrum of mass-ejection properties associated with massive stars, covering from Type II/Ib/Ic core-collapse supernovae (SNe) to giant eruptions of Luminous Blue Variables (LBV) and optical transients. In this work, we hypothesize that a large fraction of these phenomena may have an explosive origin, the distinguishing ingredient being the ratio of the prompt energy release $E_{\\rm dep}$ to the envelope binding energy $E_{\\rm binding}$. Using one-dimensional one-group radiation hydrodynamics and a set of 10-25\\,\\msun\\, massive-star models, we explore the dynamical response of a stellar envelope subject to a strong, sudden, and deeply-rooted energy release. Following energy deposition, a shock {\\it systematically} forms, crosses the progenitor envelope on a day time-scale, and breaks-out with a signal of hour-to-days duration and a 10$^5$-10$^{11}$\\,\\lsun\\, luminosity. We identify three different regimes, corresponding to a transition from dynamic to quasi-static diffusion transport. For $E_{\\rm dep} > E_{\\rm binding}$, full envelope ejection results with a SN-like bolometric luminosity and kinetic energy, modulations being commensurate to the energy deposited and echoing the diversity of Type II-Plateau SNe. For $E_{\\rm dep} \\sim E_{\\rm binding}$, partial envelope ejection results with a small expansion speed, and a more modest but year-long luminosity plateau, reminiscent of LBV eruptions or so-called SN impostors. For $E_{\\rm dep} < E_{\\rm binding}$, we obtain a ``puffed-up'' star, secularly relaxing back to thermal equilibrium. In parallel with gravitational collapse and Type II SNe, we argue that thermonuclear combustion, for example of as little as a few 0.01\\,\\msun\\, of C/O, could power a wide range of explosions/eruptions. Besides massive stars close to the Eddington limit and/or critical rotation, 8-12\\,\\msun\\, red-supergiants, which are amongst the least bound of all stars, represent attractive candidates for transient phenomena. ", "introduction": "\\label{sect_intro} Stellar explosions, broadly refered to as supernovae (SNe), are understood to stem from a sudden release of energy either associated with the collapse of the degenerate core of a massive star or from the thermonuclear combustion of fresh fuel deep inside the stellar envelope. Whether one or the other mechanism occurs seems to depend on the main-sequence mass of the progenitor star, with core collapse occuring systematically if its value is above $\\sim$8\\,\\msun\\, \\citep[hereafter WHW02]{WHW02}. Interestingly, whatever the mechanism, the typical kinetic energy of SN ejecta is on the order of 10$^{51}$\\,erg, as inferred for example for the well-studied SN 1987A (Type II peculiar; \\citealt{blinnikov_etal_2000}), for SN 1999em (Type II-Plateau, heareafter II-P; \\citealt{utrobin_07}), or for the very uniform set of events that Type Ia SNe constitutes \\citep{WKB07_Ia_lc}. The cause of this apparent degeneracy in explosion energy is, paradoxically, perhaps not so much tied to the mechanism itself, but instead to the rather uniform total envelope binding energy of the progenitor stars, on the order of 10$^{51}$\\,erg; anything falling short of that leads to a fizzle and no SN display. \\begin{figure*} \\epsfig{file=f1a.ps,width=8.5cm} \\epsfig{file=f1b.ps,width=8.5cm} \\caption{{\\it Left:} Comparison of absolute-$V$-band-magnitude light curves for an illustrative and non-exhaustive sample of SNe, SN impostors and/or erupting LBVs (violet: SN1999em, \\citealt{leonard_etal_02a,DH06_SN1999em}; blue: SN1999gi, \\citealt{leonard_etal_02b}; turquoise: SN 2005cs, \\citealt{pastorello_etal_09}; green: SN1999br, \\citealt{pastorello_etal_09}: light green: $\\eta$ Car, \\citealt{frew_04}; yellow: SN 1997bs, \\citealt{vandyk_etal_00}; red: NGC2363-V1, \\citealt{drissen_etal_01,petit_etal_06}). For each, we adopt the distance and reddening given in the associated references. The time origin is that of maximum recorded brightness. Note that all these objects have comparable effective temperatures on the order of 10,000\\,K, hence comparable bolometric correction, making the comparison of their absolute $V$-band magnitude meaningful. We also show on the left side and in black the absolute visual magnitude of the galactic red-supergiant stars studied by \\citet[black]{levesque_etal_05}, as well as the {\\it bolometric} magnitude of the O star models computed by \\citet[gray; we use the bolometric magnitude here since O stars are hot and have large bolometric corrections]{martins_etal_05}. {\\it Right:} Same as left, but now zooming in on the time of maximum brightness (the color coding is the same as in the left panel). Notice the stark contrast between SN light curves, associated with shorter/brighter events, and erupting massive stars, associated with longer/fainter events. Importantly, notice the overlap between the intrinsic brightness of $\\eta$ Car and that of the low-luminosity Type II-P SN 1999br. In this work, we propose that this diversity of radiative displays may be accomodated by a {\\it common, explosive, origin}. \\label{fig_obs_lc} } \\end{figure*} The last decade of observations of such transient phenomena has shown, however, that the radiative signatures associated with SNe (as classified in circulars) are very diverse, from very faint to very luminous, from fast-evolving to slow-evolving or fast-expanding to slow-expanding. This diversity has been observed little in Type Ia SNe, with a few peculiar events such as SN 2002ic (presence of narrow hydrogen lines in an otherwise standard Type Ia spectrum; \\citealt{hamuy_iau_02ic}) or SNLS-03D3bb (possible Type Ia SN from a super-Chandrasekhar white dwarf star; \\citealt{howell_etal_06}). In contrast, there has been a rich diversity in explosions associated (perhaps erroneously at times) with massive stars and the mechanism of core collapse. We have observed 1) Type Ic SNe, associated or not with a long-soft $\\gamma$-ray burst, and with a standard or a very large kinetic energy (SN 1998bw, \\citealt{WES_99}; SN 2002ap, \\citealt{mazzali_etal_02}); 2) a large population of Type II-Plateau (II-P) SNe in what seems to be the generic explosion of a moderate-mass red-supergiant (RSG) star (e.g. SN 2005cs, \\citealt{maund_etal_05,UC_08}); 3) a growing number of low-luminosity SNe that share properties with standard Type II-P SNe except for being significantly and globally less energetic (e.g. SN1997D, \\citealt{chugai_utrobin_00}; SN 1999br, \\citealt{pastorello_etal_04}; OT2006-1 in M85, whose status is ambiguous, see \\citealt{kulkarni_etal_2007_m85,pastorello_etal_07_m85}). We show a sample of $V$-band absolute-magnitude light curves of such core-collapse SNe in Fig.~\\ref{fig_obs_lc}, with representative peak values of $-$14 to $-$17\\,mag and a 100-day plateau duration (best seen in the right panel of that figure), hence about 6-10\\,mag brighter than their proposed RSG progenitors (shown as black crosses). From this expanded SN sample, the range of corresponding explosion energies has considerably widened, extending above and below the standard 10$^{51}$\\,erg value. Within the core-collapse SN context, this modulation is thought to stem from modulations in the energy revival of the stalled shock above the nascent proto-neutron star, in turn modulated by the stellar-core structure \\citep{burrows_etal_07a}. The stretching to low explosion energies of potential core-collapse SN events is intriguing. For the most energetic explosions belonging to points 1 and 2 above, the classification as a SN is unambiguous. However, some transient events show an ejecta/outflow kinetic energy and a peak magnitude that are SN-like, although the events did not stem from core collapse (a star is observed at that location on post-explosion/eruption images); the community calls these SN impostors (e.g. SN1997bs; Fig.~\\ref{fig_obs_lc}; \\citealt{vandyk_etal_00}). Conversely, this raises the issue whether {\\it low-energy} Type II-P SNe are associated with core collapse - they might but they need not. We illustrate this overlap in radiative properties for a sample of such objects in Fig.~\\ref{fig_obs_lc}. One such case is the Luminous Blue Variable (LBV) $\\eta$ Car, whose properties during its 1843 eruption rival those of the low-luminosity Type II-P SN1999br. $\\eta$ Car survived this gigantic eruption, which shed about 10\\,\\msun\\, of material in what now constitutes the homunculus nebula \\citep{smith_etal_2003}. In contrast to core-collapse SNe, such eruptive phenomena in massive stars have been associated with the proximity of the star to the Eddington luminosity $L_{\\rm Edd} = 4 \\pi c G M / \\kappa$ ($\\kappa$ is the mass absorption coefficient). Due to the steep dependence of luminosity to mass (e.g. with an exponent of 3.5 for main-sequence objects), this limit is easily reached by very massive stars such as $\\eta$ Car, or more generally massive blue-supergiant stars. In this context, massive stars are thought to undergo considerable mass loss when their luminosity overcomes the Eddington limit,\\footnote{Note, however, that energy will have to be supplied to the stellar envelope to push it over this limit, and in large amounts to explain such a nebula as the homunculus.} giving rise to a porosity-modulated continuum-driven outflow \\citep{shaviv_00,owocki_etal_04}. Here, this super-Eddington wind constitutes a quasi steady-state outflow, and has therefore been thought to be of a fundamentally different nature from core-collapse SN ejecta. And indeed, one refers to a wind for the former and to an ejecta for the later. This dichotomy has been exacerbated by the stark contrast in typical light curves of eruptive stars (long lived with large brightness) and core-collapse SN explosions (short lived with huge brightness). In Fig.~\\ref{fig_obs_lc}, we show two known eruptive massive stars that highlight this contrast. However, recent observations may be challenging such a strict segregation. First, the recent identification of very fast outflowing material ahead of $\\eta$ Car's homunculus now suggests that such material was accelerated by a shock, rather than driven in a quasi-steady wind, and thus connects the giant outburst to an explosive origin \\citep{smith_08_blast}. Second, the existence of interacting SNe tells us that a massive eruption can occur merely a few years before explosion. For some, e.g. SN 2006gy \\citep{smith_etal_07a,smith_etal_07b, woosley_etal_07} or SN 1994W \\citep{dessart_etal_09}, the amount is thought to be large enough to decelerate the energetic (and necessarily faster-expanding) subsequent ejection. This very strict timing of merely a few years, {\\it which is orders of magnitude smaller than evolutionary or transport time-scales}, suggests a connection between the mechanisms at the origin of the two ejections. For SN2006gy, Woosley et al. propose recurrent pair-instability pulsations, a mechanism germane to super-massive stars and therefore extremely rare. For lower mass massive stars, this short delay of a few years seems to exclude a very-long, secular, evolution for the production of the first ejection since this would have no natural timing to the comparatively instantaneous event of core collapse. Motivated by these recent observations, we explore in this paper whether this diversity of events could be reproduced by a unique and deeply-rooted mechanism, associated with the sudden energy release above the stellar core and the subsequent shock heating of the progenitor envelope. This means would communicate a large energy to the stellar envelope on a shock crossing time-scale of days rather than on a very long-diffusion time-scale of thousands of years or more. Although different in their origin, this energy release could be a weak analogue of what results in pair-instability pulsations, i.e. a nuclear flash, as identified in the 8-12\\,\\msun\\, range by \\citet{weaver_woosley_79}. In this paper, following this shock-heating hypothesis, we use 1D radiation-hydrodynamics simulations to explore the production of explosions/eruptions in stars more massive that $\\sim$10\\,\\msun\\, on the main sequence. Rather than focusing on specific models, like those potentially associated with failed supernovae \\citep{fryer_etal_09}, we parameterize the problem through a simple energy deposition, taking place with a given magnitude, over a given time, and at a given depth in a set of pre-SN progenitor star models. We do not aim at reproducing any specific observation but, through a systematic approach, try to identify important trends, in a spirit similar to that of \\citet{falk_arnett_77}. However, we depart from these authors by studying ``non-standard'' explosions. In practice, we consider cases where the energy deposited can be both smaller or larger than the binding energy of the overlying envelope, but must imperatively be released on a very short time-scale to trigger the formation of a shock. Doing so, we identify three regimes, with ``standard'' SN explosions (short-lived transients) at the high energy end, objects that we will group in the category SN ``impostors'' (long-lived transients) at intermediate energy, and variable stars at the very low energy end. Let us stress here that we do not make the claim that all massive-star eruptions, or all transients in general, stem from a strong, sudden, and deeply-rooted energy release in their envelope. Here, we make this our working hypothesis and investigate how much such an explosive scenario can explain observations. We do find that this scenario has great potential and should be considered as a possibility when examining the origin of massive-star eruptions and associated transient phenomena. \\begin{figure} \\epsfig{file=f2.ps,width=8.5cm} \\caption{Density distribution as a function of Lagrangian mass at the onset of collapse for the models of WHW02 evolved at solar metallicity. Notice the flattening density distribution above the core for increasing mass. In low-mass massive stars, the star is structured as a dense inner region (the core), and a tenuous extended H-rich envelope. \\label{fig_rho_mr} } \\end{figure} The paper is structured as follows. In \\S\\ref{sect_input}, we briefly present the stellar evolutionary models of WHW02 that we use as input for our 1D 1-group radiation-hydrodynamics simulations. We discuss the properties of the progenitor massive stars of WHW02, such as density structure and binding energy, that are relevant for the present study. We then describe in \\S\\ref{sect_model}, the numerical technique and setup for our energy deposition study. In \\S\\ref{sect_s11}, we present the results of a sequence of simulations based primarily on the 11\\,\\msun\\, model of WHW02, discussing the properties of the shocked progenitor envelope for different values of the strength (\\S\\ref{var_edep}), the depth (\\S\\ref{var_mcut}) and the duration (\\S\\ref{var_dt}) of the energy deposition. We also discuss in \\S\\ref{var_mprog} the results obtained for more massive pre-SN progenitors, ranging from 15 to 25\\,\\msun\\, on the main sequence. In \\S\\ref{sect_rad}, we present synthetic spectra computed separately with CMFGEN \\citep{HM98_lb,DH05_qs_SN,dessart_etal_09} for a few models at a representative time after shock breakout. In \\S\\ref{sect_discussion}, we discuss the implications of our results for understanding transient phenomena, and in \\S\\ref{sect_conclusion} we summarize our conclusions. ", "conclusions": "\\label{sect_conclusion} In this paper, we have presented one-dimensional one-group (gray) two-temperature radiation-hydrodynamics simulations of pre-SN massive-star envelopes subject to a sudden release of energy above their degenerate core. Although at the high energy end, the likely phenomenon is core collapse, we more generally have in mind the thermonuclear incineration of intermediate-mass elements present in shells above the core. The motivation for this study is: 1) The existence of interacting SNe (which must eject a shell at least once before core-collapse, but more likely eject multiple shells, by some means not elucidated so far); 2) the identification of fast-moving material exterior to $\\eta$ Car's homunculus (which cannot stem from radiation-driving in a wind; \\citealt{smith_08_blast}); 3) the broad range of energies inferred for Type II-P SNe, overlapping at the low-energy end with high-energy transients and massive-star eruptions. The bulk of our results stem from work on the 11\\,\\msun\\, model of WHW02, although tests with higher-mass progenitors yield the same qualitative conclusions and regimes, merely shifted quantitatively. This work is an exploration of the outcome of a strong, sudden, and deeply-rooted energy deposition taking place at the base of a massive-star envelope. We are not proposing this scenario holds for all massive-star mass ejections, but we investigate what results when this circumstance obtains. There is no doubt it does, but how frequently and how robustly is left for future study. Although this result is not new, we find that the fundamental quantity controling the outcome of a strong, sudden, and deeply-rooted energy deposition is the binding energy of the stellar envelope, which increases considerably with progenitor mass in the range 10--40\\,\\msun. What is new, however, is our study of the long-term evolution of the gas and radiative properties that result from configurations where the energy deposited is greater, on the order of, or less than the envelope binding energy. We identify three regimes, with a continuous progression from 1) SN explosions at high energy ($E_{\\rm dep} > E_{\\rm binding}$), with a complete envelope ejection associated with a 100-day long high-plateau luminosity; 2) SN impostors at the intermediate energy range ($E_{\\rm dep} \\sim E_{\\rm binding}$), with a partial envelope ejection, and a more modest but longer-lived plateau luminosity; and 3) bloated/variable stars at the low-energy end ($E_{\\rm dep} < E_{\\rm binding}$), with little or no mass ejection but with a residual puffed-up envelope of weakly-modified properties. What conditions the results is not the magnitude of the energy deposition itself but how it compares with the binding energy of the envelope exterior to the site of deposition. Hence, to achieve the same result requires more energy in a more massive progenitor star. These properties are summarized in Fig.~\\ref{fig_summary_ekin}. \\begin{figure} \\epsfig{file=f14.ps,width=8.5cm} \\caption{Correlation between various energies (normalized to the corresponding adopted energy deposition) and the energy deposition (normalized to the corresponding envelope binding energy). In black, we show the normalized asymptotic energy of the ejecta/outflow, and in blue (red) the normalized envelope/ejecta internal (kinetic) energy at the time of shock breakout. Only the models presented in Table~1 are shown here. The black curve illustrates the three regimes we have presented here, with SN explosions, SN impostors, and variable/perturbed stars as the energy deposition varies from being much larger, comparable, and smaller than the envelope binding energy. Notice how the internal energy always dominates over the kinetic energy at the time of shock breakout in the simulations performed here. \\label{fig_summary_ekin} } \\end{figure} In all simulations presented, a shock forms systematically with a strength that depends on the energy deposited. It crosses the envelope in $\\sim$1 to $\\sim$50 days, hence communicating its energy to the entire progenitor envelope {\\it quasi-instantaneously}, i.e. as opposed to, e.g., a diffusion time-scale for energy transport of 10$^4$ years or more at the corresponding depth. This shock eventually emerges at the progenitor surface with a breakout signal that varies from a duration of an hour up to a few days (modulated here by the shock speed rather than by the atmospheric-scale height), with a flux peaking at longer wavelengths for weaker shock strengths. This breakout signal is the (and may be the only) {\\it unambiguous} evidence that the subsequent ``ejection\" was triggered by shock-heating, and thus has an explosive origin. At shock breakout, the luminosity reaches a peak, then fades, before stabilizing/rising again forming an extended plateau. This plateau phase corresponds to the recombination epoch of the ejected mass, the internal energy decreasing primarily through expansion and little through radiation. It is the large stored internal energy (in the form of radiation) that keeps the ejecta optically thick and luminous (radioactive decay is neglected in our work). We find a continuum of light curves, faster-expanding ejecta being more luminous both at breakout (stronger shock) and during the plateau phase (Fig.~\\ref{fig_summary_lpeak}). The models presented in \\S\\ref{var_edep} corroborate the correlation between plateau luminosity and mid-plateau photospheric velocity identified by \\citet{hamuy_pinto_02}, and refined by \\citet{nugent_etal_06,poznanski_etal_09}. We also find that the plateau duration is anti-correlated with energy deposition (Fig.~\\ref{fig_summary_dt}). At larger energy, faster expansion leads to faster cooling and recombination so that the ejecta photosphere recedes faster in mass/radius after reaching its peak earlier. For small energy variations in this regime, interplay between kinetic and internal energy (which are comparable at breakout) yield a plateau duration that is $\\sim$100\\,d, which is on the order of Type II-P plateau lengths. For lower energy deposition, we switch slowly from a regime of dynamic diffusion to that of quasi-static diffusion. The more slowly-expanding ejecta, characterized by a slowly-decreasing optical depth with time, gives the bolometric luminosity a modest peak value and a slow evolution, with plateau durations of up to 1-2 years. The plateau phase is thus more extended and fainter for lower energy deposition, echoing the light-curve trend going from SNe to SN impostors (Fig.~\\ref{fig_obs_lc}). Note that, in our simulations, these time-scales are always 6-7 orders of magnitude larger than the time-scale for the energy deposition, which was chosen to be ten seconds in most cases. In other words, the time-scale over which the light curve evolves (basically that of radiative diffusion in an expanding medium) has no connection to the time-scale over which the energy was deposited in the first place. \\begin{figure} \\epsfig{file=f15.ps,width=8.5cm} \\caption{Correlation between the peak luminosity during the plateau phase (black dots) and the mass-weighted average ejecta velocity. We also show the correlation for the peak luminosity at shock breakout (red dots; scaled down by a factor of 1000 for convenience). Note that we include models from Tables~1 and 2. We overplot the line $L \\propto v^{1.6}$, which follows closely the distribution of points for $L_{\\rm Peak, Plateau}$ versus $\\langle v \\rangle_M$. Our radiation-hydrodynamics simulations support the correlation identified by \\citet{hamuy_pinto_02} and subsequently improved upon by \\citet{nugent_etal_06,poznanski_etal_09}. Our slope is in close agreement with that proposed in this last reference. Impressively, the relation holds over the entire domain explored. Note that there is no consideration of radioactive decay from unstable isotopes or departures from spherical symmetry, and only data points associated with the 11\\,\\msun-progenitor star are used. Relaxing these choices would likely introduce some scatter. \\label{fig_summary_lpeak} } \\end{figure} \\begin{figure} \\epsfig{file=f16.ps,width=8.5cm} \\caption{Correlation between the plateau duration (black) and the time-like quantity $R_{\\rm phot,P}/V_{\\rm phot,P}$ (ratio of the radius and the velocity at the photosphere at the time of peak-plateau brightness) versus the time-like quantity equal to the asymptotic ejecta kinetic energy divided by the peak-plateau luminosity brightness. \\label{fig_summary_dt} } \\end{figure} From our exploration and with our set of models, we find that explosions of varying strength can yield the broadest range of outcomes in {\\it low-mass} massive stars because they are characterized by a very low envelope binding energy (Fig.~\\ref{fig_eb_mr}). We indeed obtain light curves evolving from week to year time-scales (Fig.~\\ref{fig_s11_lum_all}) and ejecta expansion rates ranging from a few tens to a few thousand \\kms (Fig.~\\ref{fig_s11_phot}). An explosion/eruption producing a transient requires merely 10$^{49}$\\,erg in such objects, a value that is so low that gravitational collapse of the stellar core may not be required. And indeed, in this mass range, stellar-evolutionary calculations reveal the existence of nuclear flashes in the last nuclear-burning stages \\citep{weaver_woosley_79}, which could represent such an energy source. We therefore propose that low-mass massive stars are prime candidates for transient phenomena in the Universe, as well as prime canditates for interacting SNe such as 1994W \\citep{dessart_etal_09}. In our simulations, sudden energy deposition above the core leads to shock-heating of the {\\it entire} envelope. Whenever the energy deposited is greater than its binding energy, the {\\it entire} envelope is ejected, with an asymptotic kinetic energy that is commensurate with the energy deposited. If a subsequent energy deposition occurs (e.g. a nuclear flash followed by gravitational collapse, as needed in a fraction at least of interacting SNe), the second ejection would have low mass and little or no hydrogen. Depending on the time between the two ejections, one could get an interacting SN for a short delay (i.e. a few years) or two transients at the same location for a long delay (the first one being dim if the explosion energy is small and the second one being potentially very dim due to the low ejected mass). These scenarios are rather speculative, but they are warranted since at present most, if not all, transients have an unknown origin and are poorly characterized. In our simulations, and within the context of this work, we obtain small mass ejections only when depositing the energy close to the progenitor surface, an eventuality that seems difficult to justify physically in such massive stars. Such low-mass ejections would seem to be better suited, for example, to the surface layers of an extended white dwarf. Observationally, low-mass ejections are likely to be associated with fast transients. For transients that are both fast and faint, a low-mass ejection in a highly- or moderately-bound object seems required. At the least, our simulations for (loosely-bound) RSG stars perturbed by a small deeply-rooted energy release produce large mass ejections (the overlying hydrogen envelope) and long-faint transients. Synthetic spectra computed for a sequence of models with varying energy deposition reveal a continuous evolution from Type II-P SN-like spectra at high energy ($L\\sim$10$^8$\\,\\lsun), to low-luminosity Type II-P SN spectra at intermediate energy ($L\\sim$10$^7$\\,\\lsun), to SN-impostor-like spectra at low energy ($L\\sim$10$^6$\\,\\lsun), with, in the same order, narrower line profiles and redder/cooler spectral-energy distributions (Fig.~\\ref{fig_v1d_sed}). The results from this work should not be compromised by the approximations made in our radiation-hydrodynamics simulations. First, with one dimensionality we prevent convective transport and any structure formation through, e.g., Rayleigh-Taylor instabilities. This may alter the properties of models characterized by low expansion speeds (longer evolution time-scale); we thus plan to study this eventuality with 2D and 3D simulations in the future. Second, we deposit the energy at a large and fixed rate, independent of any feedback effects. In the case of nuclear flashes, such feedback effects could shut-off the burning prematurely. We are aware of this artifact and will attempt in future work to develop a more physically-consistent approach by investigating the conditions that may lead to shock formation, rather than assuming a setup that systematically leads to it. However, provided a shock forms, we think our results apply. Third, progenitor models may differ on a mass-by-mass comparison with other groups but the general trend of increasing binding energy with main sequence mass should hold. Fourth, one-group transport should be accurate enough since it has been shown to capture the essentials of such radiation hydrodynamics simulations \\citep{utrobin_chugai_09} - the key physics presented here takes place at large optical depth, under LTE conditions. Our finding that very modest energy perturbations can dramatically affect the structure of a massive star motivates detailed multi-dimensional hydrodynamical investigations of massive-star interiors, in particular of the last burning stages preceding core-collapse (see, e.g., \\citealt{bazan_arnett_94,bazan_arnett_98,asida_arnett_00, arnett_etal_05,meakin_arnett_06}). As shown in these hydrodynamical simulations, and more generally, we surmise that the quasi-steady state approach of standard stellar-evolutionary codes (which keep an eye on the longer-term evolution of stars) may be missing important ingredients for our understanding of massive-star evolution. This has relevance for understanding mass loss and in particular massive-star eruptions, stellar variability/stability, and interacting SNe. Such pre-SN mass ejections would also modify, and perhaps sizably, the envelope mass and structure, thereby affecting the conditions surrounding core collapse and explosion. Ongoing and forthcoming surveys/missions like Pan-STARRS, Sky Mapper, the Palomar Transient Factory, GALEX, or the Large Synoptic Survey Telescope will better reveal the diversity of explosions, eruptions, and more generally transient phenomena in the Universe. We surmise that surveys up to now have missed a large number of low-energy long-lived transients, such as low-luminosity Type II-P SNe (objects even less energetic than SN1999br) and SN impostors. It is somewhat surprising that we have not yet detected Type II SNe with Plateau durations well in excess of 100 days. Moreover, for the shock-heating solutions presented here, a breakout signal systematically takes place. At SN-like energies, the signal may be too short to be resolved \\citep{gezari_etal_08}, but for lower-energy transients, the reduced shock speed and strength would lengthen the breakout duration up to about a day, and move the peak of the spectral-energy distribution from $\\sim$100\\AA\\ to the 1000-3000\\AA\\ range. Hence, the breakout signal should be more easily detectable in such transients, allowing to distinguish between an explosive event and, e.g., a super-Eddington wind." }, "0910/0910.0261_arXiv.txt": { "abstract": "We have been analyzing a large sample of solar-like stars with and without planets in order to homogeneously measure their photospheric parameters and Carbon abundances. Our sample contains around 200 stars in the solar neighborhood observed with the ELODIE spectrograph, for which the observational data are publicly available. We performed spectral synthesis of prominent bands of C$_{2}$ and C~I lines, aiming to accurately obtain the C abundances. We intend to contribute homogeneous results to studies that compare elemental abundances in stars with and without known planets. New arguments will be brought forward to the discussion of possible chemical anomalies that have been suggested in the literature, leading us to a better understanding of the planetary formation process. In this work we focus on the C abundances in both stellar groups of our sample. ", "introduction": "The Sun is widely thought to be formed from material representative of local physical conditions in the Galaxy at the time of its formation and to represent a standard chemical composition. This {\\it homogeneity hypothesis} has been often put in question because of many improvements in the observations techniques and data analysis. With the discovery of new planetary systems, the already known {\\it heterogeneity} sources (stellar nucleosynthesis, stellar formation process) have gained a new perspective and brought new questions. It is now a fact that stars with giant planets are, on average, richer in metals than those for which no planet was detected (\\cite[Gonzalez 2006]{Gonzalez2006} and references therein). Some authors suggested that this kind of anomaly may not only involve the content of heavy elements but also some light elements like Li, C, N, and O (\\cite[Ecuvillon et al. 2006]{Ecuvillonetal2006} and references therein). Other authors found no difference in the abundance of light elements when comparing stars with and without planets (see \\cite[Ecuvillon et al. 2004]{Ecuvillonetal2004} and references therein). The studies above are not conclusive yet. We need new tests, using more accurate and homogeneous data, with a larger number of stars. Abundances of these elements in solar-like stars will bring new information that will surely help to distinguish the different stellar and planetary formation processes. This is the purpose of our work, in which we analyzed a sample of 200 stars to homogeneously measure the photospheric parameters and C abundances. ", "conclusions": "The preliminary results on C abundances are presented here for a large number of nearby solar-like stars, based on homogeneous photospheric parameters obtained from spectra with high signal-to-noise ratio and high resolution. Our analysis used public spectra from the ELODIE database, which represent about 90\\% of all data. The remaining 10\\% include many stars with detected planets and having many observations that shall be analyzed in the same way as soon as they become available to the scientific community, since they will contribute to more reliable conclusions." }, "0910/0910.0327_arXiv.txt": { "abstract": "We investigate structures of hybrid stars, which feature quark core surrounded by a hadronic matter mantle, with super-strong toroidal magnetic fields in full general relativity. Modeling the equation of state (EOS) with a first order transition by bridging the MIT bag model for the description of quark matter and the nuclear EOS by Shen et al., we numerically construct thousands of the equilibrium configurations to study the effects of the phase transition. It is found that the appearance of the quark phase can affect distributions of the magnetic fields inside the hybrid stars, making the maximum field strength about up to $30$ \\% larger than for the normal neutron stars. Using the equilibrium configurations, we explore the possible evolutionary paths to the formation of hybrid stars due to the spin-down of magnetized rotating neutron stars. We find that the energy release by the phase transition to the hybrid stars is quite large~($\\la 10^{52}~\\rm erg$) even for super strongly magnetized compact stars. Our results suggest that the strong gravitational-wave emission and the sudden spin-up signature could be observable signals of the QCD phase transition, possibly for a source out to Megaparsec distances. ", "introduction": "A very hot issue in hadronic and nuclear physics is to search the phase transition from baryons to their constitutes - deconfined quarks. Heavy ion colliders such as RHIC (Brookhaven) and LHC (CERN) are now on line to explore the QCD phase diagram for the high temperature and small baryon density regimes, for which lattice QCD calculations predict a smooth crossover to the QCD phase transition (see \\citet{stephanov04} for review). Conversely for the low temperature and high baryon density regimes, compact stars are expected to provide a unique window on the phase transition at their extreme density with super-strong magnetic field. It has been suggested long ago that quark matter may exist in the interior of compact objects \\citep{itoh70,bodmer71,witten84}. Hybrid stars (and strange quark stars) are considered to be such objects, which feature quark cores surrounded by a hadronic matter mantle (or quark cores only) (for reviews, e.g., \\citet{weber,glendenning01}). Even if the relevant conditions could be reached in a laboratory in the near future \\citep{senger}, the conditions prevailing in the compact stars are different from those produced in accelerators, i.e., the matter is long-lived, charge neutral and in $\\beta$-equilibrium with respect to weak interactions. It is therefore important to investigate the properties of such ``exotic'' stars, providing hints about the main features of matter at those extreme conditions. The formation of the quark cores in compact stars is expected to take place by a first-order phase transition \\citep{glen92,glendenning01}. Albeit still very uncertain (e.g., \\citet{horvath07}), such a transition would proceed by the conversion of initially metastable hadronic matter in the core into the new deconfined quark phase. The metastable phase could be formed as the central density of neutron stars exceeds a critical value, due to mass accretion, spin-down or cooling. Possible astrophysical cites are in the protoneutron stars during the collapse of of supernova cores (near the epoch of bounce \\citep{takahara,yasutake07} or at the late postbounce phase \\citep{gentile,nakazato08,sagert08}) and in old neutron stars accreting from their companions \\citep{benve,chau,lin06}. In whichever cases, the sudden nucleation of the exotic phase in the hadronic star will be accompanied by a core-quake and huge energy release of the gravitational binding energy. Such energy release has been proposed to explain the central engines of the gamma-ray bursts \\citep{bombaci,bere}. Possible observables of these transient phenomenon should be glitches, magnetar flares, and superbursts \\citep{alford}. In addition, the detection of gravitational waves \\citep{ioka03,yasutake07,lin06,abdikamalov08} and neutrinos \\citep{nakazato08,sagert08} generated at the moment of the phase transition, should supply us implications to unveil the mechanism of the phase transition. Here it should be noted that most of these calculations/estimations concerning the phase transition inside compact stars are limited to a non-rotating case, in which the Tolman-Oppenheimer-Volkoff equation is solved to obtain their structures \\citep{haensel86,zdunik87,muto,drago04,yasutake09}. Exceptions are for \\citet{gourgoulhon99,yasutake05, zdunik07}, in which relativistic equilibrium configurations of rotating strange stars, beyond the so-called slow rotation approximations (see references in \\citet{glendenning01}), are constructed for estimating the energy release. In \\citet{gourgoulhon99,yasutake05}, hadron matter is assumed to be converted fully to quark matter, leading to the formation of strange quark stars like in \\citet{alcock86,benvenuto89,olesen91,lugones94}. However, it is recently pointed out that the strange quark stars could be ruled out by their too fast spin-down rates via gravitational radiations from r-modes instability \\citep{madsen}. The quasi-periodic oscillations (QPOs) of strange quark stars could not be reconciled with the observations \\citep{anna}. Moreover, \\citet{mallick09} has recently claimed that magnetars cannot convert to purely quark stars, but only to hybrid stars. These suggest us to pay attention to hybrid stars rather than stars made of purely deconfined quarks. In a very recent work by \\citet{zdunik08}, the phase transition of rotating hybrid stars is discussed, however, the equation of state (EOS) is made very idealistic, assumed to model a strong phase transition, seemingly less sophisticated than the EOSs in the recent literature cited above. In addition to rotation, neutron stars observed in nature are magnetized with the typical magnetic field strength $\\sim 10^{11}$--$10^{13}$ G \\citep{Lyne}. The field strength is often much larger than the canonical value as $\\sim 10^{15}$ G for a special class of the neutron stars such as magnetars \\citep{Lattimer:2007,woods}. In a series of our recent papers \\citep{kiuchi08a,kiuchi08b,kiuchi08d}, we have studied equilibrium configurations of relativistic magnetized compact stars to apply for the understanding of their formations and evolutions, but with hadronic EOSs. Above situations motivate us to investigate the structures of relativistic hybrid stars with magnetic fields, and by using them, to estimate the energy release at the phase transition. This study is posed as an extension to the study by \\citet{yasutake05}, in which the phase transition to the rotating strange stars was investigated. Due to the unavailability of the method to construct a fully general relativistic star with arbitrarily magnetic structures (namely both with toroidal and poloidal fields), we here consider the equilibrium with purely toroidal fields as in \\citet{kiuchi08b}. It is noted that the outcomes of the recent stellar evolution calculations \\citep{Heger:2004qp} and the MHD(magnetohydrodynamics) simulations of core-collapse supernovae (\\citet{kotake04,takiwaki04,obergaulinger06,luc2007,sawai2008,kiuchi08c,takiwaki09}), suggesting much dominance of the toroidal fields, are not in contradiction with the assumption. Following the scenario proposed in \\citet{yasutake05}, we consider the possible evolutionary tracks of a rapidly rotating and magnetized neutron star to a slowing rotating hybrid star due to the spin-down via gravitational radiation and/or magnetic braking. During the evolutions, the baryon mass and the magnetic field strength are taken to be constant for simplicity. The energy release can be estimated from the difference in the mass-energies between the hadronic star and the hybrid star along each sequence. By constructing thousands of equilibrium configurations, we hope to clarify the possible maximum energy release at the moment of the transition and discuss their astrophysical implications. The paper is organized as follows. The method for constructing the EOS with the phase transition and the numerical scheme for the stellar equilibrium configurations are briefly summarized in Sec.~\\ref{sec2}. Sec.~\\ref{sec:result} is devoted to showing numerical results. Summary and discussion follow in Sec.~\\ref{sec:summary}. In this paper, we use geometrical units with $G=c=1$. ", "conclusions": "\\label{sec:summary} In this study, we investigated structures of general relativistic hybrid stars containing super-strong magnetic fields. Pushed by the results of recent stellar evolution calculations and the outcomes of recent MHD simulations of core-collapse supernovae, we treated the toroidal fields only. Using the bulk Gibbs construction, we modeled the EOS with a first order transition by bridging the MIT bag model for the description of quark matter and the nuclear EOS by Shen et al. We found that the presence of the quark phase can affect the distribution of the magnetic fields inside the hybrid stars, leading to the enhancement of the field strength about $30$ \\% than for the normal neutron stars. Using the equilibrium configurations, we explored the possible evolutionary paths to the formation of hybrid stars due to the spin-down of magnetized and rotating neutron stars. For simplicity, the total baryon mass and magnetic flux are taken to be conserved during the evolution but that the angular momenta are lost gravitational waves and/or magnetic breaking. We found that the energy release by the conversion to hybrid stars is typically $\\la 10^{52}~\\rm erg $, smaller than previously estimated for the conversion to the strange quark stars. In association with the vast energy release, it is natural to expect the emissions of gravitational waves. Amplitude of the gravitational waves associated with the phase transition, $h$, is given as follows \\citep{pacheco98}, \\begin{eqnarray} h \\sim 2 \\left( \\frac{G \\alpha E_{\\rm tot}}{c^3 \\tau} \\right)^{1/2} \\frac{1}{r \\omega}. \\end{eqnarray} Here, $E_{\\rm tot}$, $\\alpha$, $\\tau$, $r$, and $\\omega$ are energy release, fraction of the energy release emitted in form of the gravitational waves, damping time scale of the gravitational wave, distance to the source, and angular velocity, respectively. To set an absolute upper bound, we could choose $ E_{\\rm tot} \\sim 10^{52}$ erg and $\\omega \\sim 100$ rad s$^{-1}$ according to Table~\\ref{tab:Mb-const-200}. We assume that the source is in our galactic center of $r \\sim 10$ kpc and that the damping timescale and $\\alpha$ is $\\sim 100$ ms and $10^{-4}$, respectively inferred from \\citet{abdikamalov08}. The resulting amplitudes become as high as $h \\sim 10^{-18 \\sim -19}$ peaking at $\\sim$ kHz, which are surely within the detection limits for the laser interferometers on line such as LIGO, VIRGO, GEO600, and TAMA300. Considering the optimal sensitivity of the interferometers of $h \\sim 10^{-21}$ at $\\sim$ 1 KHz, such signals, far stronger than the ones from canonical core-collapse supernovae (e.g., \\citet{kotake09a,kotake09b,murphy}), are possibly visible for a source out to Megaparsec distances. In combination with the signatures of the spin-up discussed before, we speculate that the waveforms of such strong gravitational waves could provide us hints of properties of QCD phase transition. To find the accurate waveforms, general relativistic simulations are indispensable (Kiuchi et al. in preparation). Moreover accurate predictions for neutrino emissions \\citep{gentile,nakazato08,sagert08} and their detectability are remained to be studied (e.g., \\citet{kawagoe}). This paper should help to construct the initial conditions for such studies. \\citet{alcock86} claimed that the conversion from the hadronic to the mixed phase occurs instantaneously in the weak interaction timescales~(such as $ d+u \\leftrightarrow s+u$) when the central density exceeds the critical value of $\\rho_1$ in Figures 1 and 2. If so, the newly-born neutron stars may convert to the hybrid stars in the postbounce phase, before they evolve from A to B illustrated in Figure \\ref{fig:senario}. In this case, the released energy may be used to power original core-collapse supernovae or more energetic supernovae such as hypernovae. Although the formation paths to hybrid stars are different from the ones discussed in this paper, the estimation method of the energy release here are also applicable if we implement the EOS at finite temperature and high lepton fraction \\citep{yasutake09}, which we plan to study as a sequel to this paper. It has been pointed out that the super-strong magnetic field more than $10^{18}$~G can affect the stiffness of the EOS \\citep{broderick00}. As mentioned, the maximum magnetic field increases about $\\sim 30 \\%$ by the phase transition, while it increases only unit percent if the compact stars evolve to the non-rotating ones in absence of the transition. This suggests that the magnetic effects should be seriously taken into account to our models treating the phase transition. Especially for the quark phase, \\citet{fukushima07, noronha07} have recently found that the energy gaps of magnetic color-flavor-locked phase are oscillating functions of the magnetic field. Such effects are remained to be studied. We should comment about the very high values of the magnetic fields, larger than $10^{18}$G. \\citet{ruderman00} suggested that such high magnetic fields can be brought to the stellar surface by buoyancy forces. According to their result, we estimate the buoyancy time scale in our models. The buoyancy force can be expressed as \\begin{eqnarray} F_b \\sim \\frac{B_\\phi^2}{8\\pi c_s^2} g, \\label{eq:fb} \\end{eqnarray} where $c_s$ and $g$ are the sound speed and the gravitational acceleration, respectively. Putting typical values inside compact objects into the equation, we roughly estimate the time scale $\\tau$ as \\begin{eqnarray} \\tau \\sim 10^{-3}-10^{-4} {\\rm s} \\left(\\frac{\\rho}{10^{15} {\\rm g/cm^3}}\\right)^{1/2} \\left(\\frac{R}{10^6 {\\rm cm}}\\right)^{1/2} \\left(\\frac{c_s}{10^{10}{\\rm cm/s}}\\right) \\left(\\frac{B_\\phi}{10^{18} {\\rm G}}\\right)^{-1} \\left(\\frac{g}{10^{14} {\\rm cm /s^{2}}}\\right)^{-1/2}, \\end{eqnarray} where $R$ is the typical size of compact stars. Clearly such magnetic fields are not stable against the buoyancy forces in stellar evolution time scale $\\sim 10^3$yr. Based on the MHD simulations in full general relativity, \\citet{kiuchi08c} recently showed that the toroidal configurations of neutron stars is really dynamically unstable due to the buoyant instability. But more important finding relevant to this study, is that the toroidal magnetic fields settle down to a new equilibrium state with the circular motions in the meridian plane. In the new equilibrium configurations, the toroidal fields almost equivalent to the strength before the onset of the instability, are expected to be much stronger that the poloidal ones. This suggests that our models presented here, albeit unstable to the buoyant instability, could be helpful to understand equilibrium configurations of magnetized hybrid stars. Such stability analysis is an important issue yet to be studied. Our treatments of the phase transition should be more sophisticated. We plan to employ the so-called NJL model (e.g., \\citet{yasutake09}) instead of the simple MIT bag model. Moreover we shall take into account finite size effects \\citep{maruyama} such as quark-nuclear pasta structures. We considered cold compact objects, namely with zero-temperature and zero-neutrino fraction. We plan to extend this study to the finite temperature with the non-zero neutrino fraction, which should be useful for studying earlier evolutions of compact stars soon after their formation. The evolution sequence we explored is a very simplified one. For more realistic estimations, one needs the two dimensional evolutionary calculation in full general relativistic framework, not an easy job, due to the treatment of convection, combustion, cooling and/or heating processes, nucleosynthesis on the surface, and so forth. Applying the method recently reported by \\cite{pons}, we think that we could study the magneto-thermal evolution of neutron stars to the hybrid stars. By doing so, we hope to investigate the peculiar properties in the light curves, which has been pointed out to be another observationally visible \\citep{page00,blaschke00,blaschke01,grigorian05}. All of such studies should be indispensable to pave the way for the understanding of the long-veiled phase-transition physics from the astrophysical phenomena." }, "0910/0910.1117_arXiv.txt": { "abstract": "The diffusion of astrophysical magnetic fields in conducting fluids in the presence of turbulence depends on whether magnetic fields can change their topology via reconnection in highly conducting media. Recent progress in understanding fast magnetic reconnection in the presence of turbulence is reassuring that the magnetic field behavior in computer simulations and turbulent astrophysical environments is similar, as far as magnetic reconnection is concerned. This makes it meaningful to perform MHD simulations of turbulent flows in order to understand the diffusion of magnetic field in astrophysical environments. Our studies of magnetic field diffusion in turbulent medium reveal interesting new phenomena. First of all, our three-dimensional MHD simulations initiated with anti-correlating magnetic field and gaseous density exhibit at later times a de-correlation of the magnetic field and density, which corresponds well to the observations of the interstellar media. While earlier studies stressed the role of either ambipolar diffusion or time-dependent turbulent fluctuations for de-correlating magnetic field and density, we get the effect of {\\it permanent} de-correlation with one fluid code, i.e. without invoking ambipolar diffusion. In addition, in the presence of gravity and turbulence, our three-dimensional simulations show the decrease of the magnetic flux-to-mass ratio as the gaseous density at the center of the gravitational potential increases. We observe this effect both in the situations when we start with equilibrium distributions of gas and magnetic field and when we follow the evolution of collapsing dynamically unstable configurations. Thus the process of turbulent magnetic field removal should be applicable both to quasi-static subcritical molecular clouds and cores and violently collapsing supercritical entities. The increase of the gravitational potential as well as the magnetization of the gas increases the segregation of the mass and magnetic flux in the saturated final state of the simulations, supporting the notion that the reconnection-enabled diffusivity relaxes the magnetic field + gas system in the gravitational field to its minimal energy state. This effect is expected to play an important role in star formation, from its initial stages of concentrating interstellar gas to the final stages of the accretion to the forming protostar. In addition, we benchmark our codes by studying the heat transfer in magnetized compressible fluids and confirm the high rates of turbulent advection of heat obtained in an earlier study. ", "introduction": "Astrophysical flows are known to be turbulent and magnetized. The specific role played by MHD turbulence in different branches of astrophysics is still highly debated, but it is generally regarded as important. In particular, for the interstellar medium (ISM) and star formation, the role of turbulence has been discussed in many reviews (see \\citealt{elmegreen2004,mckee2007}). The opinion on the role of magnetic field in these environments vary from magnetic field being regarded as absolutely dominant in the processes (see \\citealt{tassis2005,galli2006}) to moderately important, as in super-Alfv\\'enic models of star formation (see \\citealt{padoan2004}). The vital question that frequently permeates these debates is the diffusion of the magnetic field in astrophysical fluids. The conductivity of most of the astrophysical fluids is high enough to make the Ohmic diffusion negligible on the scales involved, which means that the ``frozen-in'' approximation is a good one for many astrophysical environments. However, without considering diffusive mechanisms that can violate the flux freezing, one faces problems attempting to explain many observational facts. For example, simple estimates show that if all the magnetic flux is brought together with the material that collapses to form a star in molecular clouds, then the magnetic field in a proto-star should be several orders of magnitude higher than the one observed in T-Tauri stars (this is the ``magnetic flux problem'', see \\citealt{galli2006} and references therein, for example). To address the problem of the magnetic field diffusion both in the partially ionized ISM and in molecular clouds, researchers usually appeal to the ambipolar diffusion concept (see \\citealt{mestel1956,shu1983}). The idea of the ambipolar diffusion is very simple and may be easily exemplified in the case of gas collapsing to form a protostar. As the magnetic field is acting on charged particles only, it does not directly affect neutrals. Neutrals move under the gravitational pull but are scattered by collisions with ions and charged dust grains which are coupled with the magnetic field. The resulting flow dominated by the neutrals will be unable to drag the magnetic field lines and these will diffuse away through the infalling matter. This process of ambipolar diffusion becomes faster as the ionization ratio decreases and therefore, becomes more important in poorly ionized cloud cores. \\citet{shu2006} have explored the accretion phase in low-mass star formation and concluded that there should exist an effective diffusivity about 4 orders of magnitude larger than Ohmic diffusivity in order to an efficient magnetic flux transport to occur. They have argued that ambipolar diffusion could work, but only under rather special circumstances like, for instance, considering particular dust grain sizes. In other words, at the moment it is unclear if ambipolar diffusion is really high enough to solve the magnetic flux transport problem in collapsing flows. Does magnetic field remain absolutely frozen-in within highly ionized astrophysical fluids? The answer to this question affects the description of numerous essential processes in the interstellar and intergalactic gas. Magnetic reconnection was appealed in \\citet{lazarian2005} as a way of removing magnetic flux from gravitating clouds, e.g. from star-forming clouds. That work referred to the reconnection model of \\citet{lazarian1999} and \\citet{lazarian2004} for the justification of the concept of fast magnetic reconnection in the presence of turbulence. The advantage of the scheme proposed by \\citet{lazarian2005} was that robust removal of magnetic flux can be accomplished both in partially and fully ionized plasma, with only marginal dependence on the ionization state of the gas\\footnote{The rates were predicted to depend on the reconnection rate, which according to \\citet{lazarian2004} very weakly depends on the ionization degree of the gas.}. The concept of ``reconnection diffusion'' introduced in \\citet{lazarian2005} is relevant to our understanding of many basic astrophysical processes. In particular, it suggests that the classical textbook description of molecular clouds supported both by hourglass magnetic field and turbulence is not self-consistent. Indeed, turbulence is expected to induce ``reconnection diffusion'' which should enable fast magnetic field removal from the cloud. However, in the absence of numerical confirmation of the fast reconnection, the scheme of magnetic flux removal through ``reconnection diffusion'' as opposed to ambipolar diffusion stayed somewhat speculative. Fortunately, it has been recently shown numerically (see \\citealt{kowal2009}) that three-dimensional magnetic reconnection in turbulent fluid follows the predictions of the reconnection model of \\citet{lazarian1999} and therefore, is fast. This naturally increased the interest to the ``reconnection diffusion'' (see \\citealt{lazarian2009}). Motivated by this fact, here we perform simulations aiming to gain understanding of the diffusion of magnetic field induced by turbulence. As \\citet{kowal2009} tested magnetic reconnection in fully ionized gas in the present paper we shall focus our efforts on one fluid MHD simulations. We shall compare ``reconnection diffusion'' with the ambipolar diffusion and discuss the effects of ambipolar diffusion qualitatively focusing on the comparison of our results on ``reconnection diffusion'' with those on ``ambipolar turbulent diffusion'' described in \\citet{heitsch2004}. The latter study reported the enhancement of ambipolar diffusion in the presence of turbulence, which raised the question of how important is the simultaneous action of turbulence and ambipolar diffusion and whether turbulence alone, i.e. without any effect from ambipolar diffusion, can equally well induce de-correlation of magnetic field and density. Our work on ``reconnection diffusion'' should also be distinguished from the research on the de-correlation of magnetic field and density within compressible turbulent fluctuations. \\citet{cho_lazarian2002,cho_lazarian2003} performed three-dimensional MHD simulations and reported the existence of separate turbulent cascades of Alfv\\'en and fast modes in strongly driven turbulence as well as a cascade of slow modes driven by Alfv\\'enic cascade. Slow modes in magnetically dominated plasma are associated with density perturbations with marginal perturbation of magnetic fields, while the same is true for fast modes in weakly magnetized or high beta plasmas. Naturally, these two modes de-correlate magnetic fields and density on the crossing time of the wave. This was the effect studied in more detail in one-dimensional both analytically and numerically by \\citet{passot2003}, who stressed that the enhancements of magnetic field strength and density may correlate and anti-correlate in turbulent interstellar gas within the fluctuations and this can introduce the dispersion of the mass-to-flux ratios within the turbulent volume. Each of the fluctuations provide a {\\it transient} change of the pointwise magnetization. In the absence of other effects, e.g. related to the thermal instability, the de-correlation is reversible. In comparison, the ``reconnection diffusion'', similar to the ambipolar diffusion, deals with the {\\it permanent} de-correlation of magnetic field and density. Acting alone, the ``reconnection diffusion'' increases entropy making magnetic field-density de-correlation irreversible. What are the laws that govern \\textit{magnetic field diffusion} in turbulent magnetized fluids? Could those affect our understanding of basic interstellar and star formation processes? These are the questions that we address in this paper. In this study, we try to understand the diffusion of magnetic field in a couple of idealized models in the presence of turbulence. We explore setups both with and without gravity and compare the diffusion of magnetic field with that of a passive scalar. In the context of star formation an important issue that we will address is an alternative way of decreasing the magnetic flux-to-mass ratio without appealing to ambipolar diffusion. We claim that since turbulence in astrophysics is really ubiquitous, our results should be widely applicable. We also perform simulations including turbulent heat diffusion, which allow us to compare results obtained with our code with those in the literature. This paper is organized as follows. In Section 2, we draw the theoretical grounds about fast magnetic reconnection. In Section 3, we describe the numerical code employed. In Section 4, we present the results concerning the diffusion of magnetic field in a setup without external gravitational forces. In Section 5, we present the results of our experiments of diffusion of magnetic field in the presence of a gravitational field. In Section 6, we discuss our results and compare with previous works. In Section 7, we discuss the accomplishments and limitations of our present study. In Section 8, we discuss our findings in the context of strong turbulence theory, and finally in Section 9, we summarize our conclusions. While our work is focused on the diffusion of magnetic fields, we address in the Appendix the heat transport in magnetized turbulent plasma. We confirm with higher resolution the results in \\citet{cho2003a} that the heat advection within turbulent flows in the presence of magnetic fields is very similar to that induced by hydrodynamic turbulence. ", "conclusions": "Through this work we have performed the comparison of our results with the study by \\citet{heitsch2004}. Below, we provide yet another outlook of the connection of that study with the present paper. We also discuss the work by \\citet{shu2006}, which was the initial motivation of our study of the diffusion of magnetic field in the presence of gravity. \\subsection{Comparison with \\citet{heitsch2004}: Ambipolar Diffusion Versus Turbulence and 2.5-dimensional Versus Three-dimensional} In view of the astrophysical implications, the comparison between our results and those of \\citet{heitsch2004} calls for the discussion on how ambipolar diffusion and turbulence interact to affect the magnetic field diffusivity. In particular, \\citet{heitsch2004} claim that a new process ``turbulent ambipolar diffusion'' (see also \\citealt{zweibel2002}) acts to induce fast magnetic diffusivity. At the same time, our results do not seem to exhibit less magnetic diffusivity than those of \\citet{heitsch2004} in spite of the fact that we do not have ambipolar diffusion. How can this be understood? We propose the following explanation. In the absence of ambipolar diffusion, the turbulence propagates to smaller scales making small-scale interactions possible. On the other hand, ambipolar diffusion affects the turbulence, increasing the damping scale. As a result, the ambipolar diffusion acts in two ways, in one to increase the small-scale diffusivity of the magnetic field, in another is to decrease the turbulent small-scale diffusivity and these effects essentially compensate each other\\footnote{A possible point of confusion is related to the difference of the physical scales involved. If one associates the scale of the reconnection with the thickness of the Sweet--Parker layer, then, indeed, the ambipolar diffusion scale is much larger and therefore the reconnection scale gets irrelevant. However, within the LV99 model of reconnection, the scale of reconnection is associated with the scale of magnetic field wandering. The corresponding scale depends on the turbulent velocity and is not small.}. In other words, if we approximate the turbulent diffusivity by $(1/3) V_{\\text{inj}} L_{\\text{inj}}$, where $V_{\\text{inj}}$ and $L_{\\text{inj}}$ are the injection velocity and the injection scale for strong MHD turbulence (see LV99, \\citealt{lazarian2006}), respectively, the ambipolar diffusivity acting on small scales will not play any role and the diffusivity will be purely ``turbulent''. If, however, the ambipolar diffusion coefficient is larger than $V_{\\text{inj}} L_{\\text{inj}}$, then the Reynolds number of the steered flow may become small for strong MHD turbulence to exist and the diffusion is purely ambipolar in this case. We might speculate that this leaves little, if any, parameter space for the ``turbulent ambipolar diffusion'' when turbulence and ambipolar diffusion synergetically enhance diffusivity, acting in unison. This point should be tested by three-dimensional two-fluid simulations exhibiting both ambipolar diffusion and turbulence. In view of our findings one may ask whether it is surprising to observe the ``reconnection diffusion'' of magnetic field being independent of ambipolar diffusion. We can appeal to the fact well known in hydrodynamics, namely, that in a turbulent fluid the diffusion of a passive contaminant does not depend on the microscopic diffusivity. In the case of high microscopic diffusivity, the turbulence provides mixing down to a scale $l_1$ at which the microscopic diffusivity both, suppresses the cascade and ensures efficient diffusivity of the contaminant. In the case of low microscopic diffusivity, turbulent mixing happens down to a scale $l_2\\ll l_1$, which ensures that even low microscopic diffusivity is sufficient to provide efficient diffusion. In both cases the total effective diffusivity of the contaminant is turbulent, i.e. is given by the product of the turbulent injection scale and the turbulent velocity. This analogy is not directly applicable to ambipolar diffusion, as this is a special type of diffusion and magnetic fields are different from passive contaminants. However, we believe that our results show that to some extent the concept of turbulent diffusion developed in hydrodynamics carries over (due to fast reconnection) to magnetized fluid. \\subsection{Transient De-correlation of Density and Magnetic Field} Magneto-sonic waves are known to create transient changes of the density and magnetic field correlation. In the case of turbulence the situation is less clear, but the research in the field suggests that the decomposition of the turbulent motions into basic MHD modes is meaningful and justified even for high amplitude motions (\\citealt{cho2003a}). Thus the claim in \\citet{passot2003} that even in the limit of ideal MHD, turbulence can {\\it transiently} affect the magnetic field and density correlations is justified. However, the process discussed in this paper is different in the sense that the de-correlation we describe here is {\\it permanent} and it will not disappear if the turbulence dissipates. In a sense, as we showed above, ``reconnection diffusion'' is similar to the ambipolar and Ohmic diffusion. It is a dissipative diffusion process, which does require non-zero resistivity, although this resistivity can be infinitesimally small for the LV99 model of fast reconnection in the presence of turbulence (see Section 2). \\subsection{Relation to \\citet{shu2006}: Fast Removal of Magnetic Flux During Star Formation} As discussed in \\citet{shu2006}, the sufficiency of the ambipolar diffusion efficiency for explaining observational data of accreting proto-stars is questionable. At the same time, they found that the required dissipation is about 4 orders of magnitude larger than the expected Ohmic dissipation. Thus they appealed to the hyper-resistivity concept in order to explain the higher dissipation of magnetic field. We feel, however, that the hyper-resistivity idea is poorly justified (see criticism of it in \\citealt{lazarian2004} and \\citealt{kowal2009}). At the same time, fast three-dimensional ``reconnection diffusion'' can provide the magnetic diffusivity that is adequate for fast removing of the magnetic flux. This is what, in fact, was demonstrated in the present set of numerical simulations. It is worth mentioning that, unlike the actual Ohmic diffusivity, ``reconnection diffusion'' does not transfer the magnetic energy directly into heat. The lion share of the energy is being released in the form of kinetic energy, driving turbulence (see LV99). If the system is initially laminar, this potentially result in flares of reconnection and the corresponding diffusivity. This is in agreement with LV99 scheme where a more intensive turbulence should induce more intensive turbulent energy injection and lead to the unstable feeding of the energy of the deformed magnetic field. However, the discussion of this effect is beyond the scope of the present paper. Similar to \\citet{shu2006}, we expect to observe the heating of the media. Indeed, although we do not expect to have Ohmic heating, the kinetic energy released due to magnetic reconnection is dissipated locally and therefore we expect to observe heating in the medium. Our setup for gravity can be seen as a toy model representing the situation in \\citet{shu2006}. In the broad sense, our work confirms that a process of magnetic field diffusion that does not rely on ambipolar diffusion is efficient. We accept that our setup assuming an axial gravitational field is a very simple and ignores complications that could arise from using a nearly spherical potential of the self-gravitating cloud. The periodic boundary conditions give super-stability to the system, and do not allow inflow (or outflow) of material/magnetic field as we expect in a more realistic accretion process. However, our experiments can give us qualitative insights. They show that the turbulent diffusion of the magnetic field can remove magnetic flux from the central region, leading to a lower flux-to-mass ratio in regions of higher gravity compared with that of lower gravity. We chose parameters to the simulations such that the system is not initially unstable to the Parker--Rayleigh--Taylor (PRT) instability. Although the PRT instability could be present in real accretion systems and could help to remove magnetic field from the core of gravitational systems, its presence would make the interpretation of the results more difficult and we wanted to analyze only the turbulence role in the removal of magnetic flux. However, it is possible that this instability had been also acting due to local changes of parameters due to the turbulent motion. To ensure that the transport of magnetic flux is being caused by injection of turbulence only, we stopped the injection after a few time steps in some experiments and left the system to evolve. When we did this, the changes in the profile of the magnetic field and the other quantities stopped. We showed that the higher the strength of the gravitational force, the lower the flux-to-mass ratio is in the central region (compared with the mean value in the computational domain). This could be understood in terms of the potential energy of the system. When the gravitational potential well is deeper, more energetically favorable is the pile up of matter near the center of gravity, reducing the total potential energy of the system. When the turbulence is increased, there is an initial trend to remove more magnetic flux from the center (and consequently more inflow of matter into the center), but for the highest value of the turbulent velocity in our experiments, there is a trend to remove material (together with magnetic flux) from the center, reducing the role of the gravity, due to the fact that the gravitational energy became small compared to the kinetic energy of the system. Our results also showed that when the gas is less magnetized (higher $\\beta$, or higher values of the Alfv\\'enic Mach number $M_{A}$), the reconnection diffusion of magnetic flux is more effective, but the central flux-to-mass ratio relative to external regions is smaller for more magnetized models (low $\\beta$), compared to less magnetized models. That is, the contrast $B / \\rho$ between the inner and outer radius is higher for lower $\\beta$ (or $M_{A}$). If the turbulent diffusivity of magnetic field may explain the results in \\citet{shu2006}, one may wonder whether one can remove magnetic field by this way not only from the class of systems studied by \\citet{shu2006}, but also from less dense systems. For instance, it is frequently assumed that only ambipolar diffusion is important for the evolution of subcritical magnetized clouds \\citep{tassis2005}. Our study indicates that this conclusion may be altered in the presence of turbulence. This point, however, requires further careful study, which is beyond the scope of the present paper. In the future, we intend to study a more realistic model, e.g., with open boundary conditions and more realistic gravitational potentials. Motivated by a vital problem of the dynamics of magnetic fields in astrophysical fluids, in particular, by the magnetic flux removal in star formation, in this paper we have numerically studied the diffusion of magnetic field both in the absence and in the presence of gravitational potential. Recent work on validating the idea of LV99 model of reconnection supports our assertion that our results obtained at moderate resolution represent the dynamics of turbulent magnetic field lines in astrophysics. Our findings obtained on the basis of three-dimensional MHD numerical simulations can be briefly summarized as follows: 1. In the absence of gravitational potential the ``reconnection diffusion'' removes strong anti-correlations of magnetic field and density that we impose at the start of our simulations. The system after several turbulent eddy turnover times relaxes to a state with no clear correlation between magnetic field and density, reminiscent of the observations of the diffuse ISM by \\citet{troland1986}. 2. Our simulations that started with a quasi-static equilibrium in the presence of a gravitational potential, revealed that the turbulent diffusivity induces gas to concentrate at the center of the gravitational potential, while the magnetic field is efficiently pushed to the periphery. Thus the effect of the magnetic flux removal from collapsing clouds and cores, which is usually attributed to ambipolar diffusion effect, can be successfully accomplished without ambipolar diffusion, but in the presence of turbulence. 3. Our simulations that started in a state of dynamical collapse induced by an external gravitational potential showed that in the absence of turbulence, the flux-to-mass ratio is preserved for the collapsing gas. On the other hand, in the presence of turbulence, fast removal of magnetic field from the center of the gravitational potential occurs. This may explain the low magnetic flux-to-mass ratio observed in stars compared to the corresponding ratio of the interstellar gas. 4. As an enhanced Ohmic resistivity to remove magnetic flux from cores and accretion disks has been appealed in the literature, e.g., by \\citet{shu2006}, we have also compared models with a turbulent fluid and models without turbulence but with substantially enhanced Ohmic diffusivity. We have shown that, in terms of the magnetic flux removal, the reconnection diffusion can mimic the effect of an enhanced Ohmic resistivity. 5. In addition, our results extend earlier findings of \\citet{cho2003a} for heat advection by magnetized turbulence. We show that heat advection can be parameterized by the product of the turbulence injection scale and the turbulent velocity for a range of Alfv\\'enic and sonic Mach numbers." }, "0910/0910.3865_arXiv.txt": { "abstract": "{% The first supersoft source (SSS) identification with an optical nova in \\m31 was based on ROSAT observations. Twenty additional X-ray counterparts (mostly identified as SSS by their hardness ratios) were detected using archival ROSAT, \\xmm\\ and \\chandra\\ observations obtained before July 2002. Based on these results optical novae seem to constitute the major class of SSS in \\m31. An analysis of archival \\chandra\\ HRC-I and ACIS-I observations obtained from July 2004 to February 2005 demonstrated that \\m31 nova SSS states lasted from months to about 10 years. Several novae showed short X-ray outbursts starting within 50 d after the optical outburst and lasting only two to three months. The fraction of novae detected in soft X-rays within a year after the optical outburst was more than 30\\%. Ongoing optical nova monitoring programs, optical spectral follow-up and an up-to-date nova catalogue are essential for the X-ray work. Re-analysis of archival nova data to improve positions and find additional nova candidates are urgently needed for secure recurrent nova identifications. Dedicated \\xmm/\\chandra\\ monitoring programs for X-ray emission from optical novae covering the center area of \\m31 continue to provide interesting new results (e.g. coherent 1105 s pulsations in the SSS counterpart of nova M31N 2007-12b). The SSS light curves of novae allow us -- together with optical information -- to estimate the mass of the white dwarf, of the ejecta and the burned mass in the outburst. Observations of the central area of \\m31\\ allow us -- in contrast to observations in the Galaxy -- to monitor many novae simultaneously and proved to be prone to find many interesting SSS and nova types.} ", "introduction": "The outbursts of classical novae (CNe) are caused by explosive hydrogen burning on the white dwarf (WD) surface of a cataclysmic variable, a close binary system with transfer of material from a main sequence star to the WD. After about $10^{-7}-5\\times10^{-4}$~M$_{\\odot}$ of H-rich material are transferred to the WD, ignition under degenerate conditions takes place in the accreted envelope and a thermonuclear runaway is initiated \\citep[see e.g.][]{1998ApJ...494..680J,2005ApJ...623..398Y}. As a consequence, the envelope expands and causes the brightness of the star to increase to maximum luminosities of up to $\\sim 10^{5}$~L$_{\\odot}$. A fraction of the envelope is ejected, while a part of it remains in steady nuclear burning on the WD surface. This powers a supersoft X-ray source (SSS) which can be observed as soon as the expanding ejected envelope becomes optically thin to soft X-rays, with the spectrum of a hot ($T_{eff}: 10^{5}-10^{6}$~K) WD atmosphere \\citep[][]{1991ApJ...373L..51M}. The duration of the SSS phase is inversely related to the WD mass while the time of appearance of the SSS is determined by the mass ejected in the outburst and the ejection velocity. Models of the post-outburst WD envelope show that steady H-burning can only occur for envelope masses smaller than $\\sim 10^{-5}$~M$_{\\odot}$ \\citep[][]{2005A&A...439.1061S,1998ApJ...503..381T}, and the observed evolution of the SSS in V1974 Cyg has been successfully modeled by an envelope of $\\sim 2\\times10^{-6}$~M$_{\\odot}$ \\citep[][]{2005A&A...439.1057S}. WD envelope models show that the duration of the SSS state also depends on the metalicity of the envelope, so the monitoring of the SSS states of CNe provides constraints also on the chemical composition of the post-outburst envelope. \\citet[][]{2006ApJS..167...59H} have developed envelope and wind models that simulate the optical and X-ray light curves for several WD masses and chemical compositions. Accreting WDs in recurrent novae (RNe) are good candidates for type Ia supernovae (SNe) as RNe are believed to contain massive WDs. However, one of the main drawbacks to make RNe convincing progenitors of SNe-Ia was their low fraction in optical surveys \\citep[][]{1994ApJ...423L..31D}. In the case of CNe the ejection of material in the outburst makes it difficult to follow the long-term evolution of the WD mass. For some CNe there is a disagreement between theory and observations regarding the ejected masses, with observational determinations of the mass in the ejected shell larger than predicted by models. The duration of the SSS phase provides the only direct indicator of the post-outburst envelope mass remaining on the WD in RNe and CNe. In the case of CNe with massive WDs, the SSS state is very short ($<$100 d) and could have been easily missed in previous surveys. CNe with short SSS state are additional good candidates for SNe-Ia progenitors which makes determining their frequency very important. Nevertheless, the number and duration of SSS states observed in optical novae are small: in a systematic search of X-ray emission from CNe in the ROSAT archive \\citep[][]{2001A&A...373..542O}, found only three novae with SSS emission in X-rays from a total of 39 CNe observed less than ten years after the outburst with SSS phases lasting between 400~d and 9~yr. The \\chandra\\ and \\xmm\\ observatories have detected SSS emission for several more novae; but only a limited number of observations have been performed for each source, providing little constraints on the duration of the SSS state. Of specific interest was the monitoring of the recurrent nova RS~Oph in spring 2006 with the Swift satellite which clearly determined the end of the SSS state less than 100~d after outburst \\citep[see e.g.][]{2006ATel..838....1O} which suggests a WD mass of 1.35 M$_{\\odot}$ \\citep[][]{2006ApJ...651L.141H}. The observations of the Galactic nova V458 Vul \\citep[detected as highly variable SSS $\\sim$400 d after outburst, see e.g.][]{2008ATel.1721....1D} and nova V2491 Cyg \\citep[starting its SSS state 36~d after the optical outburst, see e.g.][]{2008ATel.1573....1N} demonstrated that each Galactic nova seems to have its own peculiarities. The small number of novae found to exhibit a SSS state, and the diversity of the duration of this state (from 10 years down to few weeks) present one of the big mysteries in the study of hydrogen burning objects over the last years. Despite an extensive target of opportunity program with \\chandra\\ and \\xmm\\ (of order 3 dozen observations during the last 4 years), little progress has been made in constraining the duration of SSS states, or to even putting constraints on the long term evolution of accreting WDs in binary systems. This now may change with the monitoring campaign for Galactic novae with the \\swift\\ satellite (see Osborne et al. in this issue). \\begin{figure} \\includegraphics[width=82mm,clip=]{fovima_2008.eps} \\caption{DSS1 image of M\\,31: the SuperLOTIS and Skinakas field of view (FoV) are overlaid as big squares. Circles show the \\chandra\\ HRC~I and \\xmm\\ EPIC PN fields. Novae with outbursts after 2000 are indicated by green dots. Big circles indicate X-ray detected novae (extracted from our \\m31\\ nova web page, see http://www.mpe.mpg.de/$\\sim$m31novae/opt/m31/M31\\_table.html).} \\label{fovima} \\end{figure} ", "conclusions": "" }, "0910/0910.1098_arXiv.txt": { "abstract": "We analyse recently acquired near-infrared {\\em Hubble} Space Telescope imaging of the GOODS-South field to search for star forming galaxies at $z\\approx 7.0$. By comparing WFC\\,3 $0.98\\mu$m $Y$-band images with ACS $z$-band ($0.85\\mu$m) images, we identify objects with colours consistent with Lyman break galaxies at z$\\simeq$6.4-7.4. This new data covers an area five times larger than that previously reported in the WFC3 imaging of the {\\em Hubble} Ultra Deep Field, and affords a valuable constraint on the bright end of the luminosity function. Using additional imaging of the region in the ACS $B$, $V$ and $i$-bands from GOODS v2.0 and the WFC3 $J$-band we attempt to remove any low-redshift interlopers. Our selection criteria yields 6 candidates brighter than $Y_{\\rm AB}=27.0$, of which all except one are detected in the ACS $z$-band imaging and are thus unlikely to be transients. Assuming all 6 candidates are at $z\\approx 7$ this implies a surface density of objects brighter than $Y_{\\rm AB}=27.0$ of 0.30$\\pm$0.12\\,arcmin$^{-2}$, a value significantly smaller than the prediction from $z\\approx 6$ luminosity function. This suggests continued evolution of the bright end of the luminosity function between $z=6\\to 7$, with number densities lower at higher redshift. ", "introduction": "In recent years our understanding of the high-redshift galaxy population has rapidly expanded with the discovery of star-forming galaxies within the first billion years ($z>5$), through the Lyman break technique using broad-band imaging (e.g. Stanway, Bunker \\& McMahon 2003; Dickinson et al.\\ 2004) and searches for Lyman-$\\alpha$ emission with narrow-band filters (e.g., Ouchi et al.\\ 2008, Ota et al.\\ 2008), and most recently gamma-ray bursts (e.g. Tanvir et al.\\ 2009, Salvaterra et al. 2009). At $z\\sim 6$, the brightest ($z<26.5$) Lyman-break ``$i$-band drop-out\" galaxies have been confirmed spectroscopically (e.g. Bunker et al.\\ 2003; Stanway et al.\\ 2004ab; Dow-Hygelund et al.\\ 2007; Vanzella et al.\\ 2009) through their Lyman-$\\alpha$ emission, confirming the validity of this photometric redshift selection. In recent weeks, the installation of the new WFC3 camera on the {\\em Hubble} Space Telescope ({\\em HST}), which has an infrared channel with a large field of view, has enabled the Lyman break technique to be pushed to $z\\sim 7-10$, revealing for the first time significant numbers of galaxy candidates close to the reionization epoch (Bunker et al.\\ 2009; Bouwens et al.\\ 2009; Oesch et al.\\ 2009; McLure et al.\\ 2009; Yan et al.\\ 2009). However, the first data to be released was the extremely deep single pointing on the Hubble Ultra Deep Field ($\\approx 4\\,{\\rm arcmin}^{2}$), reaching objects as faint at mag$_{AB}=28.5$ ($6\\,\\sigma$) in $Y$-, $J$- and $H$-bands. To probe the rare galaxies at the bright end of the luminosity function at $z\\sim 7-8$ requires a larger field of view. There have been some shallower wider area searches using ground based observations (e.g. Hickey et al. 2009, Ouchi et al. 2009, Castellano et al. 2009), but the depths probed are typically $Y_{AB}<26$ (equivalent to $L_{UV}>2\\,L^*$) and the numbers of robust candidates are small. Increasing the number of $z\\approx 7$ candidates over a wide range of magnitudes is critical to exploring the shape of the luminosity function. This is particularly important as there is strong evidence for evolution of luminosity function, with suggestions that at high redshift there is a larger relative contribution to the stellar mass and UV luminosity density from sub-$L_{*}$ systems. At $z\\sim 7$ we are close to the epoch of reionization, and an open question is the mechanism by which reionization is achieved; if we are to address the contribution of the UV from star-forming galaxies, then quantifying the luminosity function is vital (along with the escape fraction of ionizing photons from galaxies, and the hardness of the UV spectral slope). In this paper we present first results from {\\em HST}/WFC3 imaging of some of the GOODS-South field, covering an area 5 times that of the WFC3 images of the {\\em Hubble Ultra Deep Field}, and reaching $Y_{AB}=27.0$ ($6\\,\\sigma$), probing luminosities around $L^*_{UV z\\sim 6}$. In conjunction with the GOODS v2.0 Advanced Camera for Surveys (ACS) images with $B,V,i,z$ bands (Giavalisco et al.\\ 2004) we search for objects which are much brighter in the $Y$-band WFC3 filter at $1\\,\\mu$m than the $z$-band $0.9\\,\\mu$m, and are undetected at shorter wavelength. These ``$z$-drops'' are candidate $z\\sim 7$ Lyman break galaxies, and the greater area of this new dataset (compared with the Hubble Ultra Deep Field WFC3 images) means that we are likely to find brighter objects more amenable to future spectroscopic confirmation. The structure of this paper is as follows. In Section 2 the imaging data and the construction of catalogues is described. In Section 3 we describe our candidate selection and discuss the observed surface density of $z\\approx 7$ galaxies, comparing with the expected number from a range of luminosity functions. Our conclusions are presented in Section 4. Throughout we adopt the standard ``concordance'' cosmology of $\\Omega_{M}=0.3$, $\\Omega_{\\Lambda}=0.7$, and use $h_{70}=H_{0}/70 {\\rm kms^{-1}\\,Mpc^{-1}}$. All magnitudes are on the AB system (Oke \\& Gunn 1983). ", "conclusions": "In this work we have searched for star-forming galaxies at $z\\approx 7$ utilising the Lyman-break technique on newly acquired F098M $Y$-band images from WFC3 on the {\\em Hubble} Space Telescope. Through the comparison of these images to existing {\\em Hubble} ACS F850LP $z$-band images we identified objects with red colours, $(z-Y)>0.8$, indicative of a break in the spectrum. We explore an area five times larger than the recent WFC\\,3 imaging of the {\\em Hubble} Ultra Deep Field. The new wider-field data in GOODS-South probes down to $Y_{AB}=27$, equivalent to $\\approx L^*_{UV}$ at $z=7$. Using additional imaging (ACS $B$, $B$, $i$-bands from GOODS v2.0, and WFC3 F125W $J$-band) we removed contaminating objects which were either detected in the bluer ACS bands or with observed near-infrared colours inconsistent with high-redshift star forming galaxies. This selection criteria left 6 candidates down to a limiting magnitude $Y_{AB}<27.0$ (equivalent to a star formation rate of $4.5\\,{\\rm M_{\\odot}\\,yr^{-1}}$ at $z=7$) of which all but one were detected in the $z$-band (and are thus likely not to be transients). This implies a surface density of objects brighter than $Y_{\\rm AB}=27.0$ of 0.30$\\pm$0.12$\\, {\\rm arcmin}^{2}$; a value smaller than both the prediction based on the observed $z\\approx 3$ and $z\\approx 6$ luminosity functions, suggesting continued evolution of the LF beyond $z=6$. Knowledge of the surface density of dropouts in this magnitude range is crucial in constraining the luminosity function, as current estimates of the $z\\approx 7$ luminosity function indicate that an $L_{*}$ galaxy has a magnitude of $Y\\sim 27$ (e.g. Bouwens et al. 2008). Determining the UV luminosity function is crucial in order to address whether star-forming galaxies could plausibly have provided the Lyman continuum photons necessary to reionize the Universe. Given the difficulty that ground-based surveys face in probing faintward of $Y=26$, larger-area surveys with {\\em HST}/WFC3 ($\\sim 100$ arcmin$^2$) probing to similar depths ($Y(6\\sigma)\\sim 27.2$) as the dataset presented in this paper will deliver a factor of 2 improvement in the poisson uncertainty and a significant decrease in the uncertainty due to cosmic variance. \\subsection*{Acknowledgements} We would like to thank the annoyamous referee for their timely response and useful suggestions. Additionally we thank Mark Lacy for comments given during the construction of the manuscript, and Richard McMahon, Jim Dunlop, Ross McLure, Masami Ouchi, Bahram Mobasher and Michele Cirasuolo for many useful discussions about Lyman break galaxies at high redshift. Based on observations made with the NASA/ESA Hubble Space Telescope, obtained from the Data Archive at the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS 5-26555. These observations are associated with programme \\#GO/DD-11359. We are grateful to the WFC\\,3 Science Oversight Committee for making their Early Release Science observations public. SW, DS \\& ES acknowledges funding from the U.K.\\ Science and Technology Facilities Council. MJJ acknowledges the support of a RCUK fellowship. SL is supported by the Marie Curie Initial Training Network ELIXIR of the European Commission under contract PITN-GA-2008-214227." }, "0910/0910.4168_arXiv.txt": { "abstract": "We apply the axisymmetric orbit superposition modeling to estimate the mass of the supermassive black hole and dark matter halo profile of NGC~4649. We have included data sets from the Hubble Space Telescope, stellar, and globular cluster observations. Our modeling gives $\\mbh= (4.5\\pm 1.0) \\times 10^9\\Msun$ and $\\mlvobs=8.7 \\pm 1.0$ (or $\\ml_V=8.0\\pm 0.9$ after foreground Galactic extinction is corrected). We confirm the presence of a dark matter halo, but the stellar mass dominates inside the effective radius. The parameters of the dark halo are less constrained due to the sparse globular cluster data at large radii. We find that in NGC~4649 the dynamical mass profile from our modeling is consistently larger than that derived from the X-ray data over most of the radial range by roughly 60\\% to 80\\%. It implies that either some forms of non-thermal pressure need to be included, the assumed hydrostatic equilibrium may not be a good approximation in the X-ray modelings of NGC~4649, or our assumptions used in the dynamical models are biased. Our new $\\mbh$ is about two times larger than the previous published value; the earlier model did not adequately sample the orbits required to match the large tangential anisotropy in the galaxy center. If we assume that there is no dark matter, the results on the black hole mass and $\\mlvobs$ do not change significantly, which we attribute to the inclusion of {\\it HST} spectra, the sparse globular cluster kinematics, and a diffuse dark matter halo. Without the {\\it HST} data, the significance of the black hole detection is greatly reduced. ", "introduction": "\\label{sec:intro} Most nearby galaxies harbor supermassive black holes at their centers. Correlations between black hole (BH) mass and host galaxy properties \\citep{mag_etal_98,geb_etal_00,fer_mer_00,har_rix_04} have been used extensively in theoretical models in order to understand growth of the black hole and galaxy \\citep[e.g.,][]{hop_etal_08}. The latest work from Hopkins et al. suggest that a main role of a black hole is to halt star formation in the galaxy when the black hole is large enough, thereby causing the dichotomy in colors \\citep[e.g.,][]{bell_08}. While there is still an active debate as to the relative role of AGN feedback versus star formation feedback, there is a consensus that physical mechanisms for black hole growth are very important for understanding mass growth in galaxies. A concern is that the black hole correlations may have significant systematic biases, both from kinematics with poor spatial resolution and models that do not adequately include the full mass profile (see discussion in \\citealt{gul_etal_09}). Dynamical modeling of galaxies using orbit superposition offers one of the best estimates on the black hole mass \\citep[e.g.,][]{rix_etal_97,van_etal_98,cre_etal_99,geb_etal_00,geb_etal_03,val_etal_04,tho_etal_04,tho_etal_05,sio_etal_09}. Assuming axisymmetry, these models do not limit the form of the allowed velocity anisotropies. Thus, the stellar orbital structure resulting from the dynamical modeling provides a unique window on the mass growth process in the massive systems, as long as the model assumptions are valid. A particular systematic bias is shown in \\citet{geb_tho_09} where they find that the black hole mass can be underestimated in the most massive galaxies if the dark halo is not included. They find a degeneracy between the dark halo and black hole mass, since without the dark halo the stellar mass-to-light ratio is overestimated which subsequently decreases the required contribution of the black hole to constrain the central kinematics. In M87, \\citet{geb_tho_09} find that the black hole mass goes from $2.5\\times10^9$ to $6.4\\times10^9~\\Msun$ by simply running models including a dark halo. Furthermore, the uncertainties do not overlap, implying a large systematic effect. In M87, however, this degeneracy is strong since the black hole is not well resolved by the kinematic data. For galaxies with well spatially-resolved kinematics, we do not expect the degeneracy to be as significant. In addition to studying the black hole mass, there is a strong need to study the shapes of dark matter profiles. There is still little consensus for those measured in individual galaxies (e.g., PNe from \\citealt{rom_etal_03}, stellar light from \\citealt{for_geb_08}, X-rays from \\citealt{gas_etal_07,chu_etal_08,hum_etal_08,hum_etal_09}, and globular clusters from \\citealt{bri_etal_06,hwa_etal_08}). The impressive work using gravitational lensing to measure the average dark matter profiles (e.g., \\citealt{man_etal_06a,man_etal_06b}) has been able to reach out to nearly 1~Mpc. However, these results need to be compared to measurements based on individual galaxies. It is important to understand the galaxy-to-galaxy scatter in the profiles and whether there are environmental effects. This paper is part of an extensive campaign to measure both the black hole mass and the dark matter profile simultaneously, to examine the possible bias of the dark matter profile on the black hole mass estimate for galaxies with various profiles, and to compare with the gravitational potentials derived with other independent techniques such as the X-rays and weak lensing studies. Initially, we focus on the more massive galaxies, but it will be important to extend future analysis over a large mass range. \\citet{geb_tho_09} report results for M87, the most massive nearby elliptical, and in this paper we focus on NGC~4649. Both galaxies are giant ellipticals with a central surface brightness ``core''. In this paper we present the axisymmetric orbit superposition models for NGC~4649 (M60), combining data from the Hubble Space Telescope ({\\it HST}), stellar, and globular cluster observations. NGC~4649 is a giant elliptical with low surface brightness located in a subclump to the east of the main Virgo concentration \\citep{for_etal_04}. NGC~4649 has been studied extensively for its total mass profile in recent X-ray modelings \\citep[e.g.,][]{hum_etal_06,gas_etal_07,hum_etal_08}, and globular cluster studies \\citep[e.g.,][]{bri_etal_06}. The goals of our study are to place NGC~4649's black hole mass estimate on a more solid footing, to infer the properties of its dark matter halo, and more importantly to offer an independent cross-check using different dynamical tracers to the previous studies on NGC~4649. We assume a distance to NGC~4649 of 15.7 Mpc. At this distance, 1$\\arcsec$ corresponds to 76$\\;$pc. ", "conclusions": "\\label{sec:conclusions} We model the dynamical structure of NGC~4649 using the high resolution data sets from {\\it HST}, stellar, and globular cluster observations. Our main new results are: 1. Our modeling gives $\\mbh= 4.5 \\pm 1.0 \\times 10^9\\Msun$ and $\\mlvobs=8.7 \\pm 1.0$. Our new $\\mbh$ of NGC~4649 is about a factor of 2 larger than the previous result. We find that the earlier model did not adequately sample the orbits required to match the large tangential anisotropy in the galaxy center.. 2. We confirm the presence of a dark matter halo in NGC~4649, but the stellar mass dominates inside the effective radius. The parameters of the dark halo especially the core radius are less constrained due to the sparse globular cluster data at large radii. 3. Unlike in the case of M87, the black hole mass from the dynamical modeling is not biased as much by the inclusion of a dark matter halo, because high-resolution {\\it HST} spectra are available for NGC~4649, the globular cluster kinematics are sparse, and the halo is not as dominant inside the effective radius $R_e$ as that of M87. 4. We find that in NGC~4649 the dynamical mass profile from our modeling is consistently larger than that derived from the X-ray data over most of the radial range by about 70\\%. It implies that either some forms of non-thermal pressure need to be included, the assumed hydrostatic equilibrium may not be a good approximation in the X-ray modeling, or the assumptions used in our dynamical modeling create a bias." }, "0910/0910.5956.txt": { "abstract": "A study of metal enrichment of the intergalactic medium (IGM) using a series of smooth particle hydrodynamics (SPH) simulations is presented, employing models for metal cooling and the turbulent diffusion of metals and thermal energy. An adiabatic feedback mechanism was adopted where gas cooling was prevented on the timescale of supernova bubble expansion to generate galactic winds without explicit wind particles. The simulations produced a cosmic star formation history (SFH) that is broadly consistent with observations until z $\\sim$ 0.5, and a steady evolution of the universal neutral hydrogen fraction ($\\Omega_{\\rm H I}$) that compares reasonably well with observations. The evolution of the mass and metallicities in stars and various gas phases was investigated. At z=0, about 40\\% of the baryons are in the warm-hot intergalactic medium (WHIM), but most metals (80\\%-90\\%) are locked in stars. At higher redshifts the proportion of metals in the IGM is higher due to more efficient loss from galaxies. The results also indicate that IGM metals primarily reside in the WHIM throughout cosmic history, which differs from simulations with hydrodynamically decoupled explicit winds. The metallicity of the WHIM lies between 0.01 and 0.1 solar with a slight decrease at lower redshifts. The metallicity evolution of the gas inside galaxies are broadly consistent with observations, but the diffuse IGM is under enriched at z $\\sim$ 2.5. Galactic winds most efficiently enrich the IGM for halos in the intermediate mass range $10^{10}$M$_{\\sun}$ - $10^{11}$ M$_{\\sun}$. At the low mass end gas is prevented from accreting onto halos and has very low metallicities. At the high mass end, the fraction of halo baryons escaped as winds declines along with the decline of stellar mass fraction of the galaxies. This is likely because of the decrease in star formation activity and decrease in wind escape efficiency. Metals enhance cooling which allows WHIM gas to cool onto galaxies and increases star formation. Metal diffusion allows winds to mix prior to escape, decreasing the IGM metal content in favour of gas within galactic halos and star forming gas. Diffusion significantly increases the amount of gas with low metallicities and changes the density-metallicity relation. ", "introduction": "The intergalactic medium (IGM) contains most of the baryons in the Universe and it provides the fuel for galaxies to form stars in which metals are produced. In turn, supernovae and galactic winds enrich the IGM with metals, while stars and active galactic nuclei (AGN) emit UV photons. This interplay between the IGM and galaxies, mediated by metal cooling in the presence of UV, regulates the formation of stars in the universe. The evolution and enrichment history of the IGM provides a record of this interplay. Observations of metal absorption lines (e.g., C III , C IV, Si III, S IV and O VI) in quasar spectra show that the intergalactic medium (IGM) far outside large galaxies ($\\rho/\\rho_{mean} < 10$) is enriched \\citep[e.g.][]{Songaila96, Dave98, Ellison00, Schaye00, Pettini03, Schaye03, Aguirre04, Simcoe04}. There is evidence for enrichment extending back to $z > 5$ \\citep{Pettini03, Simcoe06} though metallicities do not evolve much from z=4 to z=2 \\citep{Schaye03}. Since metals are created in stars inside galaxies, those in the IGM must have escaped from galaxies. Exactly how this occurs is unclear. It typically assumed that galactic winds, driven by star formation, dominate IGM enrichment \\citep{Aguirre07}. Winds can be either launched as the ejecta of a large number of co-existing supernova (SN) explosions \\citep{Heckman90}, driven by the injection of momentum by SN and stellar winds or by radiation pressure from starbursts and AGN \\citep{Murray05}. Observations of galactic winds \\citep[e.g.,][]{Heckman01, Pettini01} have found that galactic wind velocities range from hundreds to thousands of km/s and the mass loss rates are comparable to star formation rates. Wind material has a complex, multiphase structure but most metals are expected to be entrained in the hot phase \\citep[][and references therein]{Veilleux05}. Although detailed hydrodynamical simulations of the interstellar medium (ISM) in galaxies \\citep[e.g,,][]{MacLow99, Strickland00, Williams02} have been able to generate galactic outflows and have explored various properties of winds and the gas dynamics in different phases, current cosmological simulations lack the resolution to launch or track winds directly. Hot, low density SN bubbles are unresolved in such simulations which initially led to an overcooling problem that produced unrealistically concentrated simulated galaxies \\citep{Navarro97}. As a result, various ``subgrid'' stellar feedback and wind models have emerged. These models serve two functions: to regulate star formation and the properties of the ISM and to redistribute gas (and newly formed metals) both within and into the environment around galaxies. There are three main approaches, energetic feedback, kinetic feedback and modifications to the effective equation of state which behaves similarly to an increased effective pressure. Energetic feedback in its simplest form involves simply adding the stellar feedback as thermal energy, but this suffers from overcooling \\citep{Katz96}. Kinetic feedback \\citep[e.g.,][]{Navarro93, SH03, Oppenheimer06, DallaVecchia08} converts part of the SN energy into kinetic energy in the gas. The effectiveness of kinetic feedback is strongly dependent on the resolution and hydrodynamic method. \\citet{SH03} argued that regulated star formation creates an effective pressure in the ISM and this was modelled directly in the \\textsc{gadget} code as part of a recipe for regulated star formation. This approach leads to a strongly hydrodynamically-coupled, multiphase ISM that does not naturally produce galactic outflows. To combat this the authors added a ``superwind'' model where fluid elements in the star forming region are ejected at fixed speed and are also hydrodynamically decoupled until they leaves the galaxy. \\citet{Oppenheimer06} modified the model in a manner referred to as the momentum-driven wind scenario so that the velocity of the wind and the mass loading factor were related to the velocity dispersion of the host galaxy. The \\textsc{gadget} code with superwind feedback prescriptions has been widely used in various problems such as damped Lyman-$\\alpha$ (DLA) absorbers \\citep{Nagamine04a, Nagamine04b} and the enrichment of the IGM at high and low redshifts \\citep{Oppenheimer06, Oppenheimer09}. According to these works, superwind feedback is essential to suppress overproduction of stars in galaxies and to reproduce the cosmic SFH at high redshift. It also increases the local fraction of the warm-hot intergalactic medium (WHIM) to a sufficient percentage (40\\% to 50\\%) to account for the ``missing baryons'' at z = 0. Although aspects of the model compare well with observations, some components do not. For example, the feedback may eject a large amount of cool gas from the galactic disks, which results in a low neutral hydrogen mass density $\\Omega_{\\rm H I}$ at z $<$ 2 \\citep{Nagamine04a}. Also, the interaction between winds and the ISM is usually not modeled in these simulations. \\citet{DallaVecchia08} found that the ISM plays an important role in regulating the amount of wind that escapes and the morphology of the galaxies. In their model winds are not hydrodynamically decoupled, which naturally allows for variable mass loading. A refined version of energetic feedback is adiabatic feedback which treats the overcooling problem by inhibiting gas from cooling until the hot SN bubbles can be resolved \\citep[e.g.,][]{Thacker00, Kay02, SommerLarsen03, Stinson06}. This is the approach used in this work. The pressure of the hot gas accelerates the ISM to generate winds. In the high resolution limit, this method approaches direct ISM modeling. Though energetic feedback is often referred to as supernova feedback, it can be used to model several types of stellar feedback such as winds and locally deposited radiation energy. The essential quantity is the energy injection rate as a function of the mass in stars and the current age of the stellar population. \\citet{Theuns02} used the \\citet{Kay02} adiabatic feedback model that turned off cooling for 10 Myrs for the feedback gas in their cosmological simulations. They found enough metals were carried by strong winds to produce C IV absorption lines that agreed with observations. \\citet{Aguirre05} used the same simulation to compare the optical depth of C IV and C III absorption lines from simulations with observations. The properties of the enriched IGM are not only affected by winds but also by gas cooling. Winds can enrich galactic halos and the IGM so that metal cooling significantly increases the cooling rates. \\citet{Aguirre05} found that their simulated metal enriched gas was too hot ($10^{5} \\sim 10^{7}$ K) and suggested that a lack of metal cooling was responsible for discrepancies between simulated and observed C IV absorption. \\citet{Oppenheimer06} found better agreement when they included the \\citet{Sutherland93} metal cooling model. The same model was used by \\citet{Choi09}, who investigated the effect of metal cooling on galaxy growth and found that it increases the local star forming efficiency and enhances accretion onto galaxies. However, \\citet{Sutherland93} did not include photoionization due to a ultraviolet (UV) radiation background which strongly affects the ionization states of metal species and changes the cooling rates. This is investigated in detail in the current work and also in \\citet{Wiersma09a}. Another important aspect of metal enrichment is the mixing of metals between the wind and the surrounding gas. The interstellar medium is highly turbulent and SN explosions are likely to be a major driver of the turbulence \\citep{MacLow04}. In addition, large velocity shear (such as between a wind and a gaseous halo) naturally generates turbulence and mixing. Turbulent mixing redistributes metals and thermal energy between the wind fluid and the ambient gas. This changes the metallicity, temperature and future evolution of the gas. While metal mixing is expected in strong outflows, it is still unclear how mixing impact the IGM. For example, observations by \\citet{Schaye07} found compact ($\\sim$ 100 pc), transient C IV absorbers that are highly enriched, suggesting poor chemical mixing at small scales. These absorbers were interpreted as enriched clumpy medium embedded within hot galactic wind fluids. If velocity shear is the major mechanism for turbulent mixing between winds and the surroundings, then this poor mixing could be explained if the clouds are carried by hot winds at the same speed. However to investigate this in detail, one must resolve wind structures, which is beyond current cosmological simulations. In this work we will focus on subgrid turbulent mixing models in cosmological context. In SPH simulations (which represent the majority of work in this area), the fluid is modeled by discrete particles. This implies that newly injected metals are locked into specific particles. For example, it was found that the distribution of metals from SPH simulations is too inhomogeneous compared with observations \\citep{Aguirre05}. To assess the potential importance of mixing, \\citet{Wiersma09} used SPH-smoothed metallicities and compared it with conventional particle metallicities, and found smoothing is able to generate significantly more material with low metallicities. This approach cannot capture the spread of metals over time with its impact on cooling and the thermal history of the gas. Directly modeling the turbulent ISM within a cosmological simulation is far beyond current capabilities. We employed a variant of the \\citet{Smagorinsky63} subgrid turbulent diffusion model, in which unresolved turbulent mixing is treated as a shear-dependent diffusion term. Metal cooling was calculated based on the diffused metals so that its non-local effects could be investigated. In this work, we present an analysis of a series of SPH cosmological simulations that incorporated adiabatic stellar feedback, detailed metal cooling and turbulent mixing to study the evolution and enrichment of the IGM. The feedback model was kept simple, following the adiabatic stellar feedback approach of \\citet{Stinson06}. This model has been calibrated via numerous galaxy formation studies \\citep[e.g., ][]{Governato07}. No additional wind prescriptions were used. With this approach, outflows arise from stellar feedback within the ISM and there is no distinction between the feedback that regulates star formation and that which drives galactic outflows. Thus this work establishes a baseline for the effectiveness of moderate stellar feedback coupled with key physical process absent from other work to reproduce the properties of the IGM. These results may be compared with explicit wind models. A further goal of this paper is to separately quantify the impact of metal cooling and turbulent mixing on the SFH, the global properties and the evolution of the IGM and its enrichment. In this first paper of a series, we present general results. The properties of specific metal absorbers in simulated quasar spectra will be presented in a second paper. This paper is organized as follows. Section~\\ref{method} describes the models for cosmological hydrodynamics, star formation, supernova feedback, metal cooling and metal diffusion. Section~\\ref{global} examines the cosmic SFH, global H I fraction and Ly-$\\alpha$ decrement in order to calibrate our models. Section~\\ref{mass_metal_evo} focuses on the evolution of the baryonic mass, metal fractions and metallicities in stars and different gas phases. We compare those results to the observed metal fractions and metallicities at different epochs, and with the simulations using different subgrid feedback models from \\citet{Oppenheimer06}, \\citet{Dave07} and \\citet{Wiersma09}. In section~\\ref{phasediagram} we analyze the distribution of mass and metallicity in the density-temperature phase diagram at z = 0. In section~\\ref{wind} we characterize our wind efficiency as a function of galaxy mass to obtain a better understanding of how different phases of the IGM get enriched. Where relevant, we have included detailed analysis of the effects of metal cooling and diffusion, and comparisons with observations. In the final section~\\ref{summary}, we summarize and discuss the broader implications. ", "conclusions": "\\label{summary} We investigated the enrichment of the intergalactic medium with SPH cosmological simulations using an adiabatic stellar feedback model. The simulations incorporated a self-consistent metal cooling model with an ultraviolet (UV) ionizing background along with metal diffusion that models the turbulent mixing in the IGM and the ISM. It was found that the UV background significantly alters the metal cooling rates at all temperatures from 100 K to $10^{9}$ K. Above $10^{4}$ K it decreases the cooling rate and shifts the cooling peak to higher temperature, while below $10^{4}$ K the UV increases the metal cooling rates due to the increase of free electrons. The simulations produced an SFH broadly consistent with observations to redshift z $\\sim$ 0.5, and a steady cosmic total neutral hydrogen fraction ($\\Omega_{\\rm H I}$) that compare relatively well with observations, although possible discrepancies in the observed $\\Omega_{\\rm H I}$ at z $\\sim$ 2-3 allow for more vigorous mass-loss. This demonstrates that adiabatic feedback can moderate SF while maintaining a regular supply of H I. The evolution of the mean flux decrement in the Ly-$\\alpha$ forest in our simulations is consistent with observations to z $\\sim$ 3-4, if the magnitude of the UV background is lowered with respect to the \\cite{HM05} rates in \\textsc{cloudy}. As the universe evolves, there is a rapid increase in the amount of warm-hot intergalactic medium (WHIM) and a decrease in the cooler diffuse IGM. At z=0, about 40\\% of the mass is in WHIM, consistent previous simulations with different methods \\citep{Cen06b, Oppenheimer06}. The metal content of the Universe evolves from the largest fraction being in the IGM to the majority residing in star forming gas, to ultimately being locked in stars at the present day. These trends reflect more effective wind escape at high redshift. IGM metals primarily reside in the WHIM, unlike \\citet{Oppenheimer06} and \\citet{Dave07}, whose metals largely reside in cool gas, as the wind gas starts cool with kinetic feedback and less likely to be heated up when the wind material is decoupled from the ISM. Our result is however in agreement with \\citet{Wiersma09} who also used kinetic feedback but with non-decoupled wind models. %Compared to the observations, our simulations produce a similar %estimated metal budget at z=0. At z $\\sim$ 2-3, our results are consistent %with observations for stellar and IGM (diffuse plus WHIM) metal fractions, %but have significantly more metals in the star forming gas. The mean metallicities of stars, star forming gas, galactic halo gas and the cold diffuse IGM all increase with time, but those of the WHIM and the ICM remain mostly constant with a slight decreasing trend. The metallicity of the WHIM stays between 0.01 to 0.1 $Z_{\\sun}$. The metallicity of the ICM is similar to the WHIM, which is smaller than the observed value, 0.2-0.5 $Z_{\\sun}$ \\citep{Aguirre07}, possibly due to absent AGN feedback. The metallicity evolution of the gas in galaxies compares well with DLA and sub-DLA observations from \\citet{Prochaska03} and the metallicity of the diffuse IGM at z $\\sim$ 2.5 is less than the observations from \\citet{Schaye03} and \\citet{Aguirre08}, suggesting higher resolution, or additional mechanisms for enriching the diffuse IGM are probably necessary. We characterized our galactic mass-loss and wind generation. For the current adiabatic feedback model, winds are most efficient for galaxies in the intermediate mass range of $10^{10}$M$_{\\sun}$ to about $10^{11}$M$_{\\sun}$. Below $10^{10}$M$_{\\sun}$ gas is likely to be prevented from accreting due to UV heating and remains as low-metallicity gas. Above about $10^{11}$ M$_{\\sun}$ fraction of wind gas decreases, possibly because the wind escape efficiency decreases with increasing halo potentials, or because the decline of star formation activities especially for massive halos. Most winds were hot when generated, but the ones expelled from intermediate mass range galaxies, having temperatures $\\sim 10^{5} - 10^{6}$K, could cool through metal lines and become diffuse IGM rather than WHIM. We investigated the effect of metal cooling and diffusion on the SFH, the evolution of $\\Omega_{\\rm H I}$ and the evolution of mass and metals in different phases. For metal diffusion, we further studied its effect in the density-temperature phase diagram at z =0 and the distribution of baryons as function of the galaxy mass. Metals significantly enhance the cooling of the WHIM, allowing the gas to cool and join galactic disks. Metals also enable cooling below $10^{4}$K. Metal cooling decreases the mass and metal fractions of the WHIM while increasing the metals in stars, halos and SF gas and increasing the SFR by 20\\% and $\\Omega_{\\rm H I}$ by 17\\% at z =0. With realistic diffusion included, metals mix between winds and surrounding gas before they leave the galaxies, decreasing the metal content in the WHIM and diffuse IGM but increasing it in the galactic halo and star forming gas. It prevents enriched, hot winds from creating highly-enriched low density regions, and makes the density-metallicity relation smoother so it follows a nearly log-linear relation in the density range $log(\\rho/\\rho_{mean}) = 0-4$. We performed two additional simulations with one of which having 8 times more particles, to investigate the convergence of our results and the resolution dependence of our model. Full convergence is not seen in most of our results, however the results from the high resolution run in SFH, $\\Omega_{\\rm H I}$ evolution, Ly$\\alpha$ decrement evolution are still broadly consistent with observational data, and the differences caused by resolution decrease with redshift. Also, the basic results of metal fraction evolution trend and enrichment efficiency in different gas phases remain unchanged. The impact of resolution can be seen in three aspects. Firstly, with high resolution the amount of neutral gas (H I) and SFR increase, hence the stellar mass fraction and the WHIM mass fraction related to wind activities. The difference is largest at high redshift. Secondly, the metal fraction locked up in stars increases with resolution while decreasing that in the ISM and IGM in general, suggesting a decline in wind efficiency. Thirdly, with high resolution the diffuse IGM and halos have a higher metal fraction at the cost of the WHIM metals. In future work, we will make a more detailed comparison with metal absorption line observations. We will also examine the detailed behaviour of various wind models and the effect on the immediate environment of galaxies." }, "0910/0910.5114_arXiv.txt": { "abstract": "In this article we identify and discuss various statistical and systematic effects influencing the astrometric accuracy achievable with MICADO, the near-infrared imaging camera proposed for the 42-metre European Extremely Large Telescope (E-ELT). These effects are instrumental (e.g. geometric distortion), atmospheric (e.g. chromatic differential refraction), and astronomical (reference source selection). We find that there are several phenomena having impact on $\\sim$100$\\mu$as scales, meaning they can be substantially larger than the theoretical statistical astrometric accuracy of an optical/NIR 42m-telescope. Depending on type, these effects need to be controlled via dedicated instrumental design properties or via dedicated calibration procedures. We conclude that if this is done properly, astrometric accuracies of 40$\\mu$as or better -- with 40$\\mu$as/yr in proper motions corresponding to $\\approx$20~km/s at 100~kpc distance -- can be achieved in one epoch of actual observations. ", "introduction": "The future optical/near-infrared European Extremely Large Telescope (E-ELT; see, e.g., Gilmozzi \\& Spyromilio \\cite{gilmozzi2008}), which is designed with a 42-metre aperture, will offer a substantial improvement in angular resolution compared to existing facilities. At wavelengths $\\lambda=2\\mu$m, diffraction-limited resolutions of $\\Theta\\simeq10$mas will be achieved. In terms of angular resolution in the near infrared, the E-ELT will outperform existing 8--10m-class telescopes like the VLT or Keck by factors of $\\approx$4--5 and the future James Webb Space Telescope (JWST) by factors of $\\approx$7. This increase in angular resolution should translate into a corresponding improvement in astrometric accuracy. In order to exploit the E-ELT's resolution, a German-Dutch-Italian-French consortium\\footnote{The MICADO collaboration includes: MPE Garching, Germany; USM Munich, Germany; MPIA Heidelberg, Germany; NOVA (a collaboration of the universities of Leiden, Groningen, and ASTRON Dwingeloo), The Netherlands; OAPD Padova (INAF), Italy; LESIA Paris, France} proposed the Multi-AO Imaging Camera for Deep Observations (MICADO) in February 2008. As the spatial resolution of any ground-based observatory is initially limited by the atmospheric seeing, MICADO will be equipped with a multi-conjugate adaptive optics (MCAO) system for achieving the diffraction limit of the 42m-telescope. This system uses three natural and six laser guide stars for correcting the atmospheric turbulence in a wide ($>2'$) field of view (Diolaiti et al. \\cite{diolaiti2008}). Images will be recorded by an array of 4$\\times$4 near-infrared (NIR) HAWAII-4RG detectors with 4096$\\times$4096 pixels each, covering a FOV of 53'. The instrument is sensitive to the wavelength range $0.8-2.5\\mu$m, thus covering the I, Y, J, H, K bands. For astrometric experiments the use of the data analysis software {\\sl Astro-WISE} (Valentijn et al. \\cite{valentijn2007}) is foreseen. In order to achieve its science goals (see Sect.~2 for details), MICADO needs to reach a stable (time scales of years) astrometric accuracy of approximately 50$\\mu$as. At present 8--10m class telescopes, accuracies of $\\approx$0.5\\% of a resolution element can be reached regularly (e.g. Fritz et al. \\cite{fritz2009b}). Therefore from simple scaling of results our goal \\emph{a priori} appears reasonable. However, at levels of the order of 100$\\mu$as there are several sources of statistical and systematic errors which need to be taken into account carefully. In this article we discuss those effects and analyse strategies to bypass them. We conclude that reaching an astrometric accuracy of better than 50$\\mu$as is highly challenging in terms of instrument design and data calibration but feasible. MICADO's astrometric performance should be of the same magnitude as that of the future astrometry space mission GAIA (e.g. Jordan \\cite{jordan2008}). GAIA will achieve accuracies better than $\\approx$50$\\mu$as only for bright (V$<$15.5) targets and only at the end of its mission. MICADO is expected to achieve this accuracy for targets with $K_{\\rm AB}<26$. Other space missions like SIM PlanetQuest (e.g. Edberg et al. \\cite{edberg2007}) or JASMINE (e.g. Gouda et al. \\cite{gouda2007}) also aim specifically at bright targets in order to reach accuracies of $\\approx$10$\\mu$as (at best). For illustration purposes, Fig.~\\ref{fig_eelt_gc} shows simulated observations of the nuclear star cluster of the Milky Way using both present day 8-10m class telescopes and E-ELT/MICADO. Physical parameters of the star cluster (stellar density profile, luminosity function) are taken from Genzel et al. \\cite{genzel2003}. We discuss technical details of our simulations in Sect. 4.1. These maps demonstrate the impressive progress to be expected with MICADO. \\begin{figure} \\includegraphics[height=8.8cm]{f01a.eps} \\includegraphics[height=8.8cm]{f01b.eps} \\caption{An illustration of the expected performance of the E-ELT/MICADO system. These simulated maps show the central $1''\\times1''$ (i.e. 8000AU$\\times$8000AU) of the nuclear star cluster of the Milky Way at 2.2$\\mu$m. \\emph{Top panel}: The target region as observed with present day 8-10m class telescopes. The diffraction-limited resolution is $\\approx$50mas. For comparison with actual observations, see, e.g., Genzel et al. (2003), Ghez et al. (2005). \\emph{Bottom panel}: The same field as seen by MICADO. The angular resolution is $\\approx$10mas. The improvement in detail and depth is obvious.} \\label{fig_eelt_gc} \\end{figure} Although this study is set up for the specific case of MICADO, most of its results are valid in general and therefore of interest beyond the E-ELT community. This paper is organised as follows. In Section~2, we discuss the science cases identified for MICADO. In Section~3, we review the concepts and techniques of accurate astrometry. In Section~4, we identify and analyse sources of systematic errors one by one and describe methods for minimizing those errors. We provide a summary of our results and an overall error budget in Section~5 and present our conclusions in Section~6. ", "conclusions": "In this article we have studied the capabilities expected for the NIR imager MICADO for the future 42-m European Extremely Large Telescope with respect to accurate astrometry. A variety of science cases requires long-term astrometric accuracies of $\\approx$50$\\mu$as. We discuss and quantify ten effects that potentially limit the astrometric accuracy of MICADO. We conclude that the systematic accuracy limit for astrometric observations with MICADO is $\\sigma_{\\rm sys}\\approx40\\mu$as. We find that astrometry at this accuracy level with MICADO requires the fulfillment of several conditions: \\begin{itemize} \\item All images, regardless of their distance in time, need to be combined via full coordinate transforms of second order or higher. \\item MICADO needs to be equipped with an astrometric calibration mask for monitoring the instrumental distortion. The pixel scale of the camera should not exceed the 3~mas/pix used in the current design. \\item Astrometric observations require decent integration times of at least 30 minutes per epoch. This is unavoidable in order to average out atmospheric tilt jitter. When using high-z galaxies as astrometric reference points, integration times up to about 10 hours can be necessary. \\end{itemize} It is noteworthy that the effects discussed in this article already affect observations collected with present 8m-class telescopes. In his exhaustive analysis of NIR images obtained with VLT/NACO, Fritz \\cite{fritz2009} has been able to detect signatures of chromatic differential refraction and differential tilt jitter in his astrometric dataset. He concludes that taking into account these effects can improve the accuracies down to few hundred $\\mu$as. This agrees with the findings of Lazorenko \\cite{lazorenko2006} and Lazorenko et al. \\cite{lazorenko2007} who analyze seeing-limited optical R-band ($\\lambda_{\\rm center}=655$nm) images taken with VLT/FORS1+2. They conclude that they are able to achieve astrometric precisions (but not accuracies) of $\\approx$100$\\mu$as by using a special scheme for scheduling observations and dedicated coordinate transforms (although they neglect instrumental geometric distortion). The analysis we provide here is set up for the specific case of E-ELT/MICADO, but parts of our results are valid in general. This study should thus contain valuable information for other future 30--40m telescopes. As some of the effects we discuss are actually observed in present day 8m-class telescope data, our analysis might also be helpful for the calibration of data already taken. We therefore expect that our work is of interest well beyond the E-ELT community." }, "0910/0910.3397_arXiv.txt": { "abstract": "We compare the abundances of various chemical species as derived with 3D hydrodynamical and classical 1D stellar atmosphere codes in a late-type giant characterized by \\teff$=3640$\\,K, $\\log g=1.0$, $\\MoH=0.0$. For this particular set of atmospheric parameters the 3D--1D abundance differences are generally small for neutral atoms and molecules but they may reach up to 0.3--0.4\\,dex in case of ions. The 3D--1D differences generally become increasingly more negative at higher excitation potentials and are typically largest in the optical wavelength range. Their sign can be both positive and negative, and depends on the excitation potential and wavelength of a given spectral line. While our results obtained with this particular late-type giant model suggest that 1D stellar atmosphere models may be safe to use with neutral atoms and molecules, care should be taken if they are exploited with ions. ", "introduction": " ", "conclusions": "" }, "0910/0910.4262_arXiv.txt": { "abstract": "We use photometric and spectroscopic observations of the eclipsing binary V69-47~Tuc to derive the masses, radii, and luminosities of the component stars. Based on measured systemic velocity, distance, and proper motion, the system is a member of the globular cluster 47~Tuc. The system has an orbital period of $29.5~d$ and the orbit is slightly eccentric with $e=0.056$. We obtain $M_p=0.8762\\pm 0.0048\\,M_\\odot$, $R_p=1.3148\\pm 0.0051\\,R_\\odot$, $L_p=1.94\\pm 0.21\\,L_\\odot$ for the primary and $M_s=0.8588\\pm 0.0060\\,M_\\odot$, $R_s=1.1616\\pm 0.0062\\,R_\\odot$, $L_s=1.53\\pm 0.17\\,L_\\odot$ for the secondary. These components of V69 are the first Population~II stars with masses and radii derived directly and with an accuracy of better than 1\\%. We measure an apparent distance modulus of $(m-M)_{V}=13.35\\pm 0.08$ to V69. We compare the absolute parameters of V69 with five sets of stellar evolution models and estimate the age of V69 using mass-luminosity-age, mass-radius-age, and turnoff mass - age relations. The masses, radii, and luminosities of the component stars are determined well enough that the measurement of ages is dominated by systematic differences between the evolutionary models, in particular, the adopted helium abundance. By comparing the observations to Dartmouth model isochrones we estimate the age of V69 to be 11.25$\\pm0.21$(random)$\\pm$0.85(systematic)~Gyr assuming [Fe/H]~=~-0.70, [$\\alpha$/Fe]~=~0.4, and $Y$~=~0.255. The determination of the distance to V69, and hence to 47~Tuc, can be further improved when infrared eclipse photometry is obtained for the variable. ", "introduction": "\\label{intro} Detached eclipsing double-line binary stars are the fundamental astrophysical laboratory for the determination of stellar parameters of mass and radius. Luminosities can be derived using measured parallaxes or from empirical color -- effective temperature relations. These data are the fundamental tests of stellar evolution models. Many field Population~I systems are known at solar mass and larger \\citep{andersen} and modern high accuracy measurements of masses, luminosities, and radii of the component stars are in general agreement with evolution models \\citep[see, for example,][]{lacy05,lacy08, clausen08}. Similar results are obtained for studies of individual binaries in the old open clusters NGC~188 \\citep{meibom09}, NGC~2243 \\citep{kaluzny06}, and NGC~6791 \\citep{grundahl08}. Recent wide field photometric surveys have identified numerous low mass systems, and for these K and M stars the common theme of a comparison of component properties with evolution models is that the models systematically underestimate the radii of the components in these binaries. Summaries of recent measurements can be found in \\citet{lopez06}, \\citet{lopez07}, and \\citet{blake08}. The situation is even less clear for Population~II stars. With the exception of CM Dra, for which the component masses are $\\sim$0.2$\\,M_\\odot$ \\citep{lacy77}, and the $\\omega$~Cen binary OGLEGC-V17, for which the analysis is compromised by an uncertain determination of the metallicity \\citep{thompson01,kal02}, there are no known Population~II detached double-line eclipsing binaries with main-sequence components. \\citet{torres02} used interferometric observations of HD~195987 ([Fe/H] $\\sim$ -0.6) to derive an orbit and measure the masses of the components. The radiative properties of the two components agree with a suite of models given some slight modifications of input parameters. While direct measurements of the radii of the components are not possible because HD~195987 is not an eclipsing binary, estimates of the radii can be derived from the orbital parallax, the bolometric flux, and the estimated effective temperatures. Here again the measured radii are larger than the models by some ten per cent. \\citet{boyajian08} have measured the angular diameter of the G subdwarf $\\mu$~Cas ([Fe/H $\\sim$ -0.8). This measurement provides a radius when combined with the Hipparchos parallax for $\\mu$~Cas. For this star the models underpredict this measured radius by about five per cent. The masses of the components of $\\mu$~Cas are only known to about ten per cent, so the model comparisons here are not well constrained. There is a clear need to locate and study Population~II detached eclipsing binary stars to obtain accurate masses and radii of their component stars. Stellar evolution models are becoming increasingly sophisticated and are used to fit observed cluster color-magnitude diagrams (CMD). The cluster CMD's are themselves improving in quality, with homogeneous surveys being conducted with the Hubble Space Telescope \\citep[see, for example, ][]{sarajdeni07}. Careful empirical tests of these models are an essential next step. This is the first paper in a series devoted to the study of detached eclipsing double-line binaries (DEB) in Galactic globular clusters with components on the cluster main sequence or subgiant branch. The Cluster AgeS Experiment (CASE) has the goal of determining the basic stellar parameters (masses, luminosities, and radii) of the components of cluster binaries to a precision of better than 1\\% in order to measure cluster ages and distances, and to test stellar evolution models. The methods and assumptions utilize basic and simple approaches offered by the field of eclipsing double-line spectroscopic binaries as described in \\citet{paczynski97} and \\citet{thompson01}. Previous CASE papers have discussed blue straggler systems in $\\omega$~Cen \\citep{kal07a} and 47~Tuc \\citep{kal07b}, and an SB1 binary in NGC~6397 \\cite{kal08}. The eclipsing binary V69-47~Tuc (hereinafter V69) was discovered by \\citet{weldrake} during a survey for variable stars in the field of the globular cluster 47~Tuc. They presented an $I$-band light curve for the variable and proposed an orbital period of $P=5.229$~d. The light curve phased with this period shows only one eclipse. The variable is located at the top of main sequence in the cluster color-magnitude diagram, and thus is of potential great interest for measurements of the cluster age and distance. In this paper we report the results of photometric and spectroscopic observations aimed at a determination of the absolute parameters of the components of V69. Section~\\ref{phot} describes the photometry of the variable and the determination of an orbital ephemeris. Section~\\ref{spec} presents the radial velocity observations. The combined photometric and spectroscopic element solutions are given in Section~\\ref{comb} while the membership in 47~Tuc is discussed in Section~\\ref{memb}. In Section~\\ref{age} we compare the properties of the components of V69 to a selection of stellar evolution models with an emphasis on estimating the age of the system. Finally in Section~\\ref{disc} we summarize our findings. ", "conclusions": "\\label{disc} To the best of our knowledge, these measurements of the masses and radii of the components of V69 are the first such high accuracy (better than 1\\%) measurements to be made for Population II stars. The binary is a member of the globular cluster 47~Tuc and so the determination of its distance and age applies to the cluster as well. We obtained a distance modulus of $(m-M)_{V}=13.35\\pm 0.08$. The main source of error in the distance estimate is the calibration dependent estimate of effective temperatures which we used to derive the bolometric luminosities for the component stars. A comparison of the measured masses, luminosities, and radii of the components to stellar evolution models suggests that the age of the system and hence the globular cluster 47~Tuc can be measured to a statistical accuracy of about 0.25 Gyr. However, it is important to understand the assumptions that go into any one model. In particular, the derived ages are very sensitive to the adopted helium abundance. We derive an age for 47~Tuc of 11.25$\\pm0.21$$\\pm$0.45~Gyr for a helium abundance $Y$~=~0.255 using Dartmouth model isochrones. All models give similar ages when the effects of helium abundance are taken into account. Comparison of Dartmouth evolutionary tracks calculated for the measured masses of the primary and secondary indicate that the helium abundance can be measured to an accuracy of about 0.03 for each of the components. We estimate a helium abundance of $Y = 0.269^{+0.017}_{-0.021}$ for 47~Tuc. The measured masses, radii, and luminosities of the components of V69 are consistent with Dartmouth models assuming [Fe/H] = -0.71, [$\\alpha$/Fe] = + 0.4, and Y = 0.27. The radii of both stars are known with high accuracy, and it is therefore possible to obtain a more accurate and robust distance determination based on the surface brightness method \\citep{barnes76,lacy77,thompson01}. The empirical calibration of surface brightness relations for dwarf and subgiant stars is improving \\citep{dibenedetto98,kervella,buermann06}, and it is reasonable to imagine that a distance accurate to a few per cent can be measured with accurate radii and $(V-K)$ colors. We are in the process of collecting near IR eclipse profile photometry of both V69 and V228 \\citep{kal07b} in order to measure the distances to these two binary stars in this way. These data will improve the estimates of the bolometric luminosities of the components and lead to a more accurate measurement of the helium abundance and hence the absolute age of 47 Tuc. Finally, we note that contributions to the errors in the radii are dominated by the photometric solution. Given the large inclination, the errors in the masses are completely dominated by the orbital solution. An identical doubling of the existing set of radial velocity observations leads to a 33$\\%$ improvement in the mass estimates, and a subsequent improvement in the age estimates. The system is relatively bright, and the prospects are good that a substantial improvement in the measured masses can be achieved with further radial velocity observations." }, "0910/0910.1453_arXiv.txt": { "abstract": "We present an analysis of timing residual (noise) of $54$ pulsars obtained from 25-m radio telescope at Urumqi Observatory with a time span of $5\\sim 8$ years, dealing with statistics of the Hurst parameter. The majority of these pulsars were selected to have timing noise that look like white noise rather than smooth curves. The results are compared with artificial series of different constant pairwise covariances. Despite the noise like appearance, many timing residual series showed Hurst parameters significantly deviated from that of independent series. We concluded that Hurst parameter may be capable of detecting dependence in timing residual and of distinguishing chaotic behavior from random processes. ", "introduction": "All pulsars show a remarkable uniformity of rotation rate on a time scale of a few days as expected of an isolated spinning body with large stable moment of inertia. \\citep{Lyne06} The angular momentum of radio pulsar is slowly decreasing through slowdown torque of the magnetic dipole radiation. However, some very interesting irregularities in pulsar rotation have been observed which are termed as {\\em timing noise}. It is anticipated that valuable information of many interesting physical processes related to pulsars is coded in the timing noises, thus employing statistical measures to characterize timing noises is important to the study of pulsars and thus the properties of the matter at supra-nuclear densities. Efforts to quantify timing noise have been tried as early as timing noise was firstly recognized, for instance according to random walk of different quantities \\citep{Boynton72}, most current models such as vortex creeping are still restricted to treatment of timing noise only as random process in certain quantities. One exception was presented by \\citep{Harding90}, who analyzed timing data of Vela pulsar to look for evidence of chaotic behavior by ``correlation sum'' technique to estimate fractal dimension of the system. However, despite possible suggestions they concluded that ``correlation sum'' estimator may be unable to distinguish between random and chaotic processes. However, any statistical representation of data has their own biases, employing a large types of statistical measures is essentially vital to a fair understanding of the timing noises. Furthermore, the number of observed pulsars has accumulated to $\\sim 10^3$ \\citep{Manchester06} but collected data is often incomplete for a conclusive analysis, it is therefore crucial to diagnose current available but limited data, the results of which could guide us in future observation to concentrate on those pulsars with anomalous timing noise. Here we are introducing a statistical method rarely used in time domain astronomy -- the Hurst parameter analysis, which is actually sensitive to the type of the inherent correlation among the time series. Our practical analysis of the timing data observed by the 25-meter radio telescope at Urumqi Observatory of $54$ pulsars indicates that Hurst parameter analysis might be capable of detecting anomalous signals which disguise themselves as noises. ", "conclusions": "\\label{sec_cncl} We calculated Hurst parameter of $54$ radio pulsars with white-noise-like timing residual obtained from Nanshan telescope and compared the results with artificial (anti-)persistent series. The majority of pulsars from our selection have Hurst parameters around $0.5$ and not far from Hurst parameter calculated for independent series. However, we found 9 pulsars (PSRs J0147$+$5922, J0357$+$5236, J0612$+$3721, J0630$-$2834, J0823$+$0159 and J0837$-$4135 showing persistent trend and PSRs J0055$+$5117, J1022$+$1001, and J1842$-$0359 showing anti-persistent trend) with interesting $H$ values despite having white-noise-like timing residual. Comparison with artificial series confirm that these trends cannot be attributed entirely to effects caused by ROS algorithm or large uncertainty in timing residual. This shows that our algorithm may be capable of detecting hidden correlation in apparently noise-like timing residual. We therefore suggest that these 9 pulsars be monitored continuously to confirm or disprove long-range dependence and search for possible physical process behind such correlation. As for the selection, we picked $54$ pulsars with relatively small period derivative $\\dot{P}$, and white-noise-like timing residual rather than the typical red-noise pattern of the three kind of noise model (PN, FN and SN) first proposed by Boynton et al. (1972) and of course we made no attempt at doing correlation between Hurst parameter and other pulsar properties nor obtaining statistics of Hurst parameter values for large population. Our method is aimed at finding possibility that timing noise that resembles white-noise is not really generated by a random process. As is well known there are pulsars with smooth timing noise that differ largely from white noise (such as J0332+5434, J0406+6138, J0826+2637, etc.) and they would yields quite large Hurst parameter if calculated using our algorithm and some even exceed $1$. In that case we do not need the deviation from $0.5$ to conclude the obvious dependent nature of timing noise and other method should be used to analyze possible chaotic nature of these timing noise series. There are various physical processes that might be responsible for the long-range dependence of pulsar timing noise. These can be classified into three groups: from interior of neutron star, i.e. due to fluctuation of internal (e.g. micro-quake due to partial release of elastic energy \\cite{Pines72} and random pinning and unpinning of vortex lines \\cite{Packard72}, \\cite{Anderson75}) and external (e.g. accretion flow \\cite{Lamb78}) torques; from emission process (e.g. magnetosphere activities \\cite{Cheng87}) and from propagation of radio emission. For the last class of origin, it has long been proposed that timing noise can be used to set upper limits on gravitational wave (\\cite{Bertotti83}) and that given enough time pulsar timing array might be the first equipment to directly detect gravitational wave (\\cite{Manchester062}). All of these (including gravitational wave detection which aims at detecting random background) predict randomness in evolution of certain quantity while Hurst parameter is capable of uncovering chaotic behavior that is hidden in timing noise. The physical origin that leads to dependence series is much more limited than random fluctuations. Therefore, long-term dependence detected by Hurst parameter might reveal more detailed information about the physical origin of timing noise. For instance it has been proposed that under certain conditions Euler equations for rotating object with magnetic dipole moment misaligned with rotation axis would show chaotic spin-down behavior (see \\cite{Harding90}). By definition any dynamical process can exhibit chaotic behavior and thus dependence in long-term timing noise data if the system is governed by differential equations whose solution is highly sensitive to initial conditions. We expect that the application of Hurst parameter to timing noise of longer time span reveal more evidence for new physics." }, "0910/0910.4937_arXiv.txt": { "abstract": "{Sub-millimeter and Far-IR observations have shown the presence of a significant amount of warm (few hundred K) and dense ($n(\\rm H_2)\\ge10^4~\\3cm$) gas in sources ranging from active star forming regions to the vicinity of the Galactic center. Since the main cooling lines of the gas phase are important tracers of the interstellar medium in Galactic and extragalactic sources, proper and detailed understanding of their emission, and the ambient conditions of the emitting gas, is necessary for a robust interpretation of the observations. } {With high resolution ($7''-9''$) maps ($\\sim3\\times3$ pc$^2$) of mid-$J$ molecular lines we aim to probe the physical conditions and spatial distribution of the warm (50 to few hundred K) and dense gas ($n(\\rm H_2)>10^5~\\3cm$) across the interface region of the nearly edge-on M17 SW nebula. } {We have used the dual color multiple pixel receiver CHAMP$^+$ on APEX telescope to obtain a $5'.3\\times4'.7$ map of the $J=6\\rightarrow5$ and $J=7\\rightarrow6$ transitions of \\twco, the \\thco $J=6\\rightarrow5$ line, and the $^3P_2\\rightarrow{^3P_1}$ 370 $\\mu$m fine-structure transition of \\ci in M17 SW. LTE and non-LTE radiative transfer models are used to constrain the ambient conditions. } {The warm gas extends up to a distance of $\\sim2.2$ pc from the M17 SW ridge. The \\thco\\ $J=6\\rightarrow5$ and \\ci\\ 370 $\\mu$m lines have a narrower spatial extent of about 1.3 pc along a strip line at P.A=$63^{\\circ}$. The structure and distribution of the \\ci $^3P_2\\rightarrow{^3P_1}$ 370 $\\mu$m map indicate that its emission arises from the interclump medium with densities of the order of $10^3~\\3cm$. } {The warmest gas is located along the ridge of the cloud, close to the ionization front. An LTE approximation indicates that the excitation temperature of the embedded clumps goes up to $\\sim120$ K. The non-LTE model suggests that the kinetic temperature at four selected positions cannot exceed 230 K in clumps of density $n(\\rm H_2)\\sim5\\times10^5~\\3cm$, and that the warm ($T_k>100$ K) and dense ($n(\\rm H_2)\\ge10^4~\\3cm$) gas traced by the mid-$J$ \\twco lines represent just about 2\\% of the bulk of the molecular gas. The clump volume filling factor ranges between 0.04 and 0.11 at these positions. } ", "introduction": "The heating and cooling balance in photon-dominated regions (PDRs) remains an active study of research. The comprehensive understanding of PDRs requires observations of large areas close to radiation sources, and of a wide wavelength range covering various emissions of atoms, molecules, and grains. In particular, mid-$J$ CO lines have been detected in almost all known massive Galactic star forming regions (e.g. Orion Nebula, W51, Cepheus A, NGC 2024). This indicates that warm ($T_K\\ge50$ K) and dense ($n(H_2)\\ge10^4~\\3cm$) gas is common, and probably of importance in most OB star forming regions. The mid-$J$ CO lines detected in regions like, e.g. M17, Cepheus A and W51, have relatively narrow line widths of 5--10 \\kms, although not as narrow as the line widths observed in cold quiescent cloud cores. Observations of the $J=6\\rightarrow5$ and $J=7\\rightarrow6$ transitions of \\twco in several massive star forming regions indicate that the warm emitting gas is confined to narrow ($<1$ pc) zones close to the ionization front. These observations favor photoelectric heating of the warm gas by UV radiation fields outside the HII regions (e.g. Harris \\etal\\ 1987; Graf \\etal\\ 1993; Yamamoto \\etal\\ 2001; Kramer \\etal\\ 2004 and 2008). Nevertheless, shocks may also be an important source of heating in high velocity wing sources like, Orion, W51 and W49 (Jaffe \\etal\\ 1987). Because of its nearly edge-on geometry, and the large amount of observational data available in the literature, M17 SW is one of the best Galactic regions to study the entire structure of PDRs from the exciting sources to the ionization front, and the succession (or not) of H$_2$, \\ci and CO emissions, as predicted by PDR models (Icke, Gatley, \\& Israel 1980; Felli, Churchwell, \\& Massi 1984; Meixner et al. 1992; Meijerink \\& Spaans 2005). M17 SW is also one of the few star forming regions for which the magnetic field strength can be measured in the PDR interface, and where the structure of the neutral and molecular gas seems to be dominated by magnetic pressure rather than by gas pressure (Pellegrini \\etal\\ 2007). M17 SW is a giant molecular cloud at a distance of 2.2 kpc, illuminated by a highly obscured ($A_v>10$ mag) cluster of several OB stars (among $\\ga100$ stars) at about 1 pc to the East (Beetz \\etal\\ 1976; Hanson \\etal\\ 1997). It also harbours a number of candidate young stellar objects that have recently been found (Povich \\etal\\ 2009). Several studies of molecular emission, excitation and line profiles (e.g. Snell \\etal~1984; Martin, Sanders \\& Hills 1984; Stutzki \\& G\\\"usten 1990) from the M17 SW core indicate that the structure of the gas is highly clumped rather than homogeneous. Emission of \\ci and \\cii was detected more than a parsec into the molecular cloud along cuts through the interface region (Keene \\etal\\ 1985; Genzel \\etal\\ 1988; Stutzki \\etal\\ 1988). These results, as well as those found in other star forming regions like S106, the Orion Molecular Cloud, and the NGC~7023 Nebula (e.g. Gerin \\& Phillips 1998; Yamamoto \\etal\\ 2001; Schneider \\etal\\ 2002, 2003; Mookerjea \\etal\\ 2003) do not agree with the atomic and molecular stratification predicted by standard steady-state PDR models. However, the extended \\ci $^3P_1\\rightarrow{^3P_0}$ and \\thco $J=2\\rightarrow1$ emission in S140 have been successfully explained by a stationary, but clumpy, PDR model (Spaans 1996; Spaans \\& van Dishoeck 1997). Hence, the lack of stratification in \\ci, \\cii and CO is a result that can be expected for inhomogeneous clouds, where each clump acts as an individual PDR. On the other hand, a partial face-on illumination of the molecular clouds would also suppress stratification. Based on analysis of low-$J$ lines of \\twco, \\thco and CH$_3$CCH data, the temperature towards the M17 SW cloud core has been estimated as 50--60 K, whereas the mean cloud temperature has been found to be about 30--35 K (e.g. G\\\"usten \\& Fiebig 1988; Bergin \\etal\\ 1994; Wilson \\etal\\ 1999; Howe \\etal\\ 2000; Snell \\etal\\ 2000). Temperatures of $\\sim275$ K has been estimated from NH$_3$ observations (G\\\"usten \\& Fiebig 1988) towards the VLA continuum arc, which agree with estimates from highly excited \\twco transitions (Harris \\etal\\ 1987). Multitransition CS and HC$_3$N observations indicates that the density at the core region of M17 SW is about $6\\times10^5~\\3cm$ (e.g., Snell \\etal\\ 1984; Wang \\etal\\ 1993; Bergin, Snell \\& Goldsmith 1996). While densities up to $3\\times10^6~\\3cm$ have been estimated towards the north rim with multitransition observations of NH$_3$, which indicates that ammonia is coexistent with high density material traced in CS and HCN (G\\\"usten \\& Fiebig 1988). The UV radiation field $G_0$ has been estimated to be of the order of $10^4$ in units of the ambient interstellar radiation field ($1.2\\times10^{-4}~\\rm ergs~s^{-1}~cm^{-1}~sr^{-1}$, Habing 1968; Meixner \\etal\\ 1992). However, most of the millimeter-wave molecular observations in M17 SW are sensitive only to low temperatures ($<100$ K), and the few available data of mid-$J$ CO and \\ci lines (consisting mostly of cuts across the ionization front and observations at few selected positions) are limited in spatial resolution and extent (e.g. Harris \\etal\\ 1987; Stutzki \\etal\\ 1988; Genzel \\etal\\ 1988; Stutzki \\& G\\\"usten 1990; Meixner \\etal\\ 1992; Graf \\etal\\ 1993; Howe \\etal\\ 2000). Therefore, in this work we present maps ($\\sim3\\times3$ pc$^2$) of mid-$J$ molecular (\\twco and \\thco) and atomic (\\ci) gas, with excellent high resolution ($9.4''-7.7''$), which advances existing work in M17 SW. The observations were done with CHAMP$^+$ (Carbon Heterodyne Array of the MPIfR) on the Atacama Pathfinder EXperiment (APEX\\footnote{This publication is based on data acquired with the Atacama Pathfinder Experiment (APEX). APEX is a collaboration between the Max-Planck-Institut fur Radioastronomie, the European Southern Observatory, and the Onsala Space Observatory}) (G\\\"usten \\etal\\ 2006). The multiple pixels at two submm frequencies of CHAMP+, allow for efficient mapping of $\\sim$arcmin regions, and provide the ability to observe simultaneously the emission from the $J=6\\rightarrow5$ and $J=7\\rightarrow6$ rotational transitions of \\twco at 691.473 GHz and 806.652 GHz, respectively. We also observed the $J=6\\rightarrow5$ transition of \\thco at 661.067 GHz and the $^3P_2\\rightarrow~^3P_1$ 370 $\\mu$m (hereafter: $2\\rightarrow1$) fine-structure transition of \\ci at 809.342 GHz. Since the gas phase cools mainly via the atomic fine structure lines of \\oi, \\cii, \\ci, and the rotational CO lines (e.g. Kaufman \\etal\\ 1999, Meijerink \\& Spaans 2005), these carbon bearing species presented here are very important coolants in the interstellar medium (ISM) of a variety of sources in the Universe, from Galactic star forming regions, the Milky Way as a galaxy, and external galaxies up to high redshifts (e.g. Fixsen \\etal\\ 1999; Weiss \\etal\\ 2003; Kramer \\etal\\ 2005; Bayet \\etal\\ 2006; Jakob \\etal\\ 2007). The case of M17 SW can be considered as a proxy for extra galactic star forming regions. M17 SW is not special, nor does it need to be, compared to other massive star-forming regions like Orion, W49, Cepheus A, or W51. Still, it does allow feedback effects, expected to be important for starburst and active galaxies, to be studied in great spatial detail. Comparison of the local line ratios to extra-galactic regions can then shed light on the properties of massive star forming regions that drive the energetics of active galaxies. Our results will be of great use for future high resolution observations, since molecular clouds of the size of the maps we present will be resolved by ALMA at the distance ($\\sim14$ Mpc) of galaxies like NGC~1068. The main purpose of this work is to explore the actual spatial distribution of the mid-$J$ \\twco\\ and \\ci lines in M17 SW, and to test the ambient conditions of the warm gas. A simple LTE model based on the ratio between the \\twco and \\thco $J=6\\rightarrow5$ lines is used to probe the temperature of the warm ($T_K\\sim100$ K) and dense ($n_{\\rm H}>10^5~\\3cm$) molecular gas. Then a non-LTE model is used to test the ambient conditions at four selected positions. In a follow up work we will present an elaborate model of these high resolution data. The most frequent references to Stutzki \\etal\\ (1988), Stutzki \\& G\\\"usten (1990) and Meixner \\etal\\ (1992) will be referred to as S88, SG90 and M92, respectively. The organization of this article is as follows. In \\S2 we describe the observations. The maps of the four lines observed are presented in \\S3. The modelling and analysis of the ambient conditions are presented in \\S4. And the conclusions and final remarks are presented in \\S5. ", "conclusions": "We have used the dual color heterodyne receiver array of 7 pixels CHAMP$^+$ on the APEX telescope, to map a region of about 2.6 pc $\\times$ 2.9 pc in the $J = 6\\rightarrow5$ and $J = 7\\rightarrow6$ lines of \\twco, the \\thco $J = 6\\rightarrow5$ and the $^3P_2\\rightarrow{^3P_1}$ 370 $\\mu$m ($J=2\\rightarrow1$) fine-structure transition of \\ci in M17 SW nebula. The completely different structure and distribution of the $^3P_2\\rightarrow{^3P_1}$ 370 $\\mu$m emission, and its critical density, indicate that this emission arises from the interclump medium ($\\sim3\\times10^3~\\3cm$). On the other hand, the mid-$J$ lines of \\twco and the isotope emissions, arise from the high density ($\\sim5\\times10^5~\\3cm$) and clumpy region. The spatial extent of the warm gas (40-230 K) traced by the \\twco $J=7\\rightarrow6$ line is about 2.2 pc from the ridge of the M17 SW complex, which is smaller than the extent observed in the low-$J$ \\twco and C$^{18}$O lines reported in previous work. The \\thco\\ $J=6\\rightarrow5$ and \\ci\\ 370 $\\mu$m lines, have a narrower spatial extent of about 1.3 pc along a strip line at P.A=$63^{\\circ}$. An LTE approximation of the excitation temperature provides lower limits for the kinetic temperature. The warmest gas is located along the ridge of the cloud, close to the ionization front. In this region the excitation temperature range between 40 and 120 K. A non-LTE estimate of the ambient conditions at four selected positions of M17 SW indicates that the high density clumps ($\\sim5\\times10^5~\\3cm$) cannot have temperatures higher than 230 K. The warm ($T_k>100$ K) and dense ($n(\\rm H_2)\\ge10^4~\\3cm$) gas traced at the four selected positions by the mid-$J$ \\twco lines represent $\\sim2$\\% of the bulk of the molecular gas traced by the low-$J$ \\twco lines. Volume filling factors of the warm gas ranging from 0.04 to 0.11 were found at these positions." }, "0910/0910.4330_arXiv.txt": { "abstract": "% There is no shortage of energy around to solve the overcooling problem of cooling flow clusters. AGNs, as well as gravitational energy are both energetic enough to balance the cooling of cores of clusters. The challenge is to couple this energy to the baryons efficiently enough, and to distribute the energy in a manner that will not contradict observational constraints of metalicity and entropy profiles. Here we propose that if a small fraction of the baryons that are accreted to the cluster halo are in the form of cold clumps, they would interact with the hot gas component via hydrodynamic drag. We show that such clumps carry enough energy, penetrate to the center, and heat the core significantly. We then study the dynamic response of the cluster to this kind of heating using a 1D hydrodynamic simulation with sub-grid clump heating, and produce reasonable entropy profile in a dynamic self-consistent way. ", "introduction": "Galaxy clusters grow by accreting dark matter and baryons from their surroundings. This is partly a continuous process of relatively smooth accretion, and partly via mergers. The existence of a visible, extended X-ray halo as well as limits on the size of the CD galaxy are two indicators that the smooth, continuous accretion is the more pronounced of the two. The baryons that are accreted in this process settle in hydrostatic equilibrium within the cluster's potential well. During this process they achieve virial equilibrium by passing through the virial shock and converting the kinetic energy of the infall to thermal energy. The baryons, at a temperature of a few $keVs$, cool primary by emitting Bremsstrahlung radiation which easily escape the halo and is observed by X-ray telescopes. After de-projection (tomography) of the luminosity and spectrum, radial profiles of temperature and densities are derived, and one can deduce the cooling times of the gas, which, for the centers of cooling flow cluster, is $\\le 1Gyr$, much shorter than the age of typical clusters that, according to $\\Lambda CDM$ formation history \\citep{press74} should have been in place at $z \\ge 1$. Had this gas cooled, \\emph{i)} cool gas would have been seen in the halo (no gas below $T=\\frac{1}{3}T_{vir}$ is observed), \\emph{ii)} a census of all the baryons in galaxies is significantly smaller than amount of gas that was expected to cool from the halo, and \\emph{iii)} the star formation within the CD would have been $10^2-10^3M_\\odot/yr$ (two orders of magnitude larger than typically observed). These three contradictions are three manifestations of the overcooling problem of cooling flow clusters. It is highly unlikely that there is a ``hidden'' baryonic component in cluster halos, so most explanations invoke some kind of heating mechanisms that would balance the cooling, keeping the gas hot and diffuse. The energies needed to compensate for the rapid cooling of cluster halos are $\\sim 10^{45}erg/sec$ which, over the lifetime of the cluster amounts to $\\sim 10^{62} erg$. These required energy rates can originate from AGN emission from the CD \\citep[][as well as many others]{ciotti97}, and by gravitational energy. Diffuse baryons falling into a gravitational well convert gravitational energy to kinetic, and ultimately to thermal. This thermal input into the system is usually local, and acts to heat the infalling gas itself. It is necessary to couple this freshly accreted, hot gas, with the central cooling gas. \\citet{narayan01} have studied conduction and turbulence that could potentially couple the external part to the inner halo. \\citet{kim03}, as well as others, deduce that although the amount of energy is sufficient, the conduction coefficient is not enough and turbulence will produce the wrong entropy and metalicity profiles. In this proceedings, we shall present a novel mechanism of gravitational heating of the central halo gas by the hydrodynamic interactions between cold clumps of accreted gas and the halo gas \\citep{db08} and \\citetext{Birnboim \\& Dekel 2009, in preparation}. The structure of this proceeding paper is as follows: First (section \\ref{sec:energetics}) we show that the accreting gas carries a sufficient rate of energy to compensate for the cooling. Then we model infalling cold clumps, and study the physical processes of their interaction with the diffuse gas, instabilities and survivability (section \\ref{sec:physics}). Using this model, we study the valid parameter space of these clumps (section \\ref{sec:monte}). We further use these insights to construct a sub-grid model for 1D hydrodynamic simulation and present simulations of a cluster with and without such clumps (section \\ref{sec:hydro}). Finally we discuss possible origins of these clumps, and summarize (section \\ref{sec:summary}). ", "conclusions": "\\label{sec:summary} We have shown that a sufficient amount of energy is released in the gravitational accretion to solve the overcooling problem of clusters (larger than $M_{vir}\\ge 10^{13}M_\\odot$), if we assume that some part of the baryons penetrates to the inner parts rather that being stopped at the virial radius. The coupling between the incoming accreted mass and the ambient gas is achieved by assuming that the accreted baryons includes gaseous cold clump, that penetrate through the virial shock and heat the gas via hydrodynamic drag. A parameter space survey indicates that for halos larger than $10^{13}M_\\odot$ and clump masses in the range $10^5\\le m_c \\le 10^8M_\\odot$, clump heating can potentially solve the overcooling problem. The dynamic response of cluster halos to clump heating, cooling and mass deposition in the inner parts in then examined. For the set of values $M_{vir}=3 \\cdot 10^{14}M_\\odot$, $m_c=10^7M_\\odot$, $f_b+f_c=0.1$, $f_c=0.01$, the resulting entropy and temperature profiles match typical observed cooling flow clusters reasonably well. The physical processes discussed here are already included in 3D hydrodynamic simulations but we note that to simulate the drag and the formation of such clumps each $10^7M_\\odot$ clump need to be well resolved, which is typically not the case for $10^{15}M_\\odot$ clusters." }, "0910/0910.5336_arXiv.txt": { "abstract": "The X-3.4 class flare of 13 December 2006 was observed with a high cadence of 2 minutes at 0.2 arc-sec resolution by HINODE/SOT FG instrument. The flare ribbons could be seen in G-band images also. A careful analysis of these observations after proper registration of images show flare related changes in penumbral filaments of the associated sunspot, for the first time. The observations of sunspot deformation, decay of penumbral area and changes in magnetic flux during large flares have been reported earlier in the literature. In this {\\it Letter}, we report lateral motion of the penumbral filaments in a sheared region of the $\\delta$-sunspot during the X-class flare. Such shifts have not been seen earlier. The lateral motion occurs in two phases, (i) motion before the flare ribbons move across the penumbral filaments and (ii) motion afterwards. The former motion is directed away from expanding flare ribbons and lasts for about four minutes. The latter motion is directed in the opposite direction and lasts for more than forty minutes. Further, we locate a patch in adjacent opposite polarity spot moving in opposite direction to the penumbral filaments. Together these patches represent conjugate foot-points on either side of the polarity inversion line (PIL), moving towards each other. This converging motion could be interpreted as shrinkage of field lines. ", "introduction": "Photospheric changes during large flares have been studied for a long time \\citep{Giovanelli40,Howard63,Rust76,Tanaka78}. Such changes have been observed successfully in various parameters like, shear angle \\citep{Ambastha93,Wang94,Hagyard99}, sunspot morphology \\citep{Anwar93}, magnetic flux \\citep{Lara00, Spirock02, Wang02a, Wang02b, Sudol05, Zharkova2005}, penumbral size \\citep{Wang04,Deng05, Liu05}. Most of these observed changes are located in the penumbra of the sunspot. \\cite{Hagyard99} argued that penumbral regions contribute most to magnetic energy of a force-free field (Low 1985) given by $E_M= \\int{(x B_x + y B_y)B_z dx dy}$. Since $B_z=0$ near Polarity Inversion Line (PIL) and $B_T = \\sqrt{B_x^2 + B_y^2} \\approx 0$ in the umbra, these regions contribute little to magnetic energy of the force-free field. Since, flares are driven by magnetic energy changes, the most promising locations to look for flare related changes are the regions between umbra and PIL. This could explain why most of the flare related studies show changes in the penumbra. \\cite{Hagyard99} found interesting coherent variation of the shear angle in patches of the magnetic field on timescales shorter than the photospheric Alfven travel time over the patches. They had conjectured that electrodynamic effects of coronal magnetic field variations could induce coherent magnetic variations even at the photospheric levels. They also concluded that instruments with greater sensitivity could throw more light on these induced field variations. The Solar Optical Telescope (SOT; \\cite{Tsuneta08}) onboard Japanese space mission {\\it Hinode} \\citep{Kosugi07} is one such instrument capable of detecting these changes. The X-3.4 class flare, observed in NOAA 10930 on 13 December 2006 during 02:00 UT to 03:00 UT, by {\\it Hinode} spacecraft, provided such an opportunity. Various aspects of this active region have been reported. Three dimensional magnetic field reconstruction of this region was performed by \\cite{Schrijver08} using NLFFF extrapolation methods. They found regions of strong vertical current density between the two spots which decreased significantly after the flare. Further, the free energy also decreased after the flare. The variation of the weighted mean shear and total magnetic shear \\citep{Wang08} with altitude was studied using extrapolated fields by \\cite{Jing08}. They found that near the PIL, up to a height of $\\approx 8 Mm$ the shear increases after the flare and decreases beyond that up to $\\approx 70 Mm$. Beyond 70 Mm the field remains close to potential before and after the flare. However, the cadence of these vector magnetograms is insufficient to search for similar flare related changes as observed by \\cite{Hagyard99}. In order to detect flare related changes in magnetized regions like sunspots one can also use high angular resolution images obtained at a high-cadence, during the flare interval. Such observations are nowadays routinely available from Filtergraph (FG; \\cite{Tsuneta08}) instrument onboard {\\it Hinode} space mission. The high resolution images allow us to investigate the evolution of the fine structure like penumbral filaments during a flare, e.g lateral motion and/or twist. In this {\\it Letter}, we use FG observations from {\\it Hinode} to investigate flare-induced changes during a X-class flare. In section 2.1 we describe the observations while the method of registration applied on the data to establish the coherent motions are described in section 2.2. We present our results in section 3 while the discussion and conclusions are given in section 4. ", "conclusions": "In this {\\it Letter}, we present sub-arcsec (0.1\"/pixel) observations in G-band at a cadence of 2 minutes, taken during 13 December 2006 X3.4 flare. High-resolution observations of the photosphere during a large X-class flare have never been done before with the kind of stability and image quality that HINODE/SOT provides. A careful analysis, after proper registration of these images, show flare related changes in penumbral filaments of the associated sunspot, for the first time, in the form of coherent lateral motions of the penumbral filaments. We found the motion of the penumbral filaments to exist in two phases: (i) the first phase is short-lived (about 4 minutes) with the time of maximum displacement coinciding with the peak of the microwave flux, and (ii) the second phase is of longer duration (more than 40 minutes). Further, we observe the patches corresponding to conjugate foot-points in either spots moving towards each other. These conjugate foot-points, labeled as `1' and `2' and marked by arrows, are clearly seen in the online G-band movie. This converging motion of conjugate foot-points could be interpreted as shrinkage of field lines during flare, which is conjectured by \\cite{Hudson00}. However, we lack high-cadence vector field measurements in this case and our present inferences are purely from G-band images. We have also shown, in section 3.4, a decrease in the global twist after the flare. It will be interesting to see the evolution of local twist as well as global twist during the flare. We expect to detect flare-related vector field changes with high-cadence vector magnetograms from the upcoming SDO mission. \\cite{Schrijver08} found strong arching filamentary current system embedded in the magnetic flux emerging between the penumbrae of the two spots, prior to flare. These currents are significantly reduced after the flare. This current system seems co-spatial with the field lines that connect the conjugate foot-points discussed above. Currents in astrophysical plasmas are produced by the distortions of the magnetic fields given by the curl of the field \\citep{Parker79}. Generally these distortions are driven by plasma dynamics. The coherent nature of the distortions in the present observations, occurring within a time scale shorter than the Alfven travel time over the affected area, rules out the plasma dynamics of the photosphere as the cause of the distortions. It is more likely that the distortions are caused by the fields induced in the photosphere in response to the restructuring (like untwisting of the global field, as inferred from the decrease in SSA) of the coronal magnetic fields, which are line tied to the photosphere. We have estimated the energy of the induced lateral motions of the penumbral filaments to be $\\sim 1.45 \\times 10^{30}$ ergs, which is a small fraction of the total magnetic energy of $ \\sim 3 \\times 10^{32}$ ergs released in the flare \\citep{Schrijver08}. This mechanism of induced photospheric magnetic response to coronal magnetic restructuring appears energetically feasible to drive the lateral motion of the filaments. We call this first phase the ``rapid restructuring phase\" (RRP). The slow global change in the displacement of the filaments during second phase can then be considered as a response of the gradual restructuring of the field on the much longer Alfven time-scale corresponding to the larger size of the CME flux rope system \\citep{Liu2008}. We call this phase the ``gradual restructuring phase\" (GRP). However, we need more observations to confirm these phases in other flares. Vector magnetic field maps at a cadence similar to G-band images are not available from SOT/SP. However, future instruments, like HMI aboard SDO mission, will have high-resolution vector magnetic field maps at high-cadence and will shed more light on the flare related vector field changes associated with such lateral motions of small-scale features." }, "0910/0910.4383_arXiv.txt": { "abstract": "We use our most recent training set for the {\\sc Rico} code to estimate the impact of recombination uncertainties on the posterior probability distributions which will be obtained from future CMB experiments, and in particular the {\\sc Planck} satellite. Using a Monte Carlo Markov Chain analysis to sample the posterior distribution of the cosmological parameters, we find that {\\sc Planck} will have biases of $-0.7$, $-0.3$ and $-0.4$ sigmas for $\\nS$, $\\omegab$ and $\\logAs$, respectively, in the minimal six parameter $\\Lambda$CDM model, if the description of the recombination history given by {\\sc Rico} is not used. The remaining parameters (e.g. $\\tau$ or $\\omegadm$) are not significantly affected. We also show, that the cosmology dependence of the corrections to the recombination history modeled with {\\sc Rico} has a negligible impact on the posterior distributions obtained for the case of the {\\sc Planck} satellite. In practice, this implies that the inclusion of additional corrections to existing recombination codes can be achieved using simple cosmology-independent `fudge functions'. Finally, we also investigated the impact of some recent improvements in the treatment of hydrogen recombination which are still not included in the current version of our training set for {\\sc Rico}, by assuming that the cosmology dependence of those corrections can be neglected. In summary, with our current understanding of the complete recombination process, the expected biases in the cosmological parameters inferred from {\\sc Planck} might be as large as $-2.3$, $-1.7$ and $-1$ sigmas for $\\nS$, $\\omegab$ and $\\logAs$ respectively, if all those corrections are not taken into account. We note that although the list of physical processes that could be of importance for {\\sc Planck} seems to be nearly complete, still some effort has to be put in the validation of the results obtained by the different groups. The new {\\sc Rico} training set as well as the fudge functions used for this paper are publicly availabe in the {\\sc Rico}-webpage. ", "introduction": "The cosmic microwave background (CMB) is nowadays an essential tool of theoretical and observational cosmology. Recent advances in the observations of the CMB angular fluctuations in temperature and polarization \\citep[e.g. ][]{WMAP5-basic,WMAP5-params} provide a detailed description of the global properties of the Universe, and the cosmological parameters are currently known with accuracies of the order of few percent in many cases. The experimental prospect for the {\\sc Planck} satellite \\citep{Planck2006}, which was launched on May 14th 2009, is to achieve the most detailed picture of the CMB anisotropies down to angular scales of $\\ell \\sim 2500$ in temperature and $\\ell \\sim 1500$ in polarization. This data will achieve sub-percent precision in many cosmological parameters. However, those high accuracies will rely on a highly precise description of the theoretical predictions for the different cosmological models. Currently, it is widely recognised that the major limiting factor in the accuracy of angular power spectrum calculations is the uncertainty in the ionization history of the Universe \\citep[see ][]{hu1995, Seljak2003}. This has motivated several groups to re-examine the problem of cosmological recombination \\citep{Zeldovich1968, Peebles1968}, taking into account detailed corrections to the physical processes occurring during hydrogen \\citep[e.g.][]{Dubrovich2005, RHS2005, Chluba2006, Kholupenko2006, Rubino2006, Chluba2007, Chluba2007a, Karshenboim2008, Hirata2008, Chluba2009c, Chluba2009b, Chluba2009, Jentschura2009, Labzowsky2009, HirataForbes2009} and helium recombination \\citep[e.g][]{Kholupenko2007, Wong2007, Switzer2008a, Switzer2008b, HirataSwi2008, Rubino2008, Kholupenko2008, Chluba2009d}. Each one of the aforementioned corrections individually leads to changes in the ionization history at the level of $\\ga 0.1$\\%, in such a way that the corresponding overall uncertainty in the CMB angular power spectra exceeds the benchmark of $\\pm 3/\\ell$ at large $\\ell$ \\citep[for more details, see][hereafter FCRW09]{rico}, thus biasing any parameter constraints inferred by experiments like {\\sc Planck}, which will be cosmic variance limited up to very high multipoles. The standard description of the recombination process is provided by the widely used {\\sc Recfast} code \\citep{Seager1999}, which uses effective three-level atoms, both for hydrogen and helium, with the inclusion of a conveniently chosen {\\it fudge factor} which artificially modifies the dynamics of the process to reproduce the results of a multilevel recombination code \\citep{Seager2000}. The simultaneous evaluation of all the new effects discussed above make the numerical computations very time-consuming, as they currently require the solution of the full multilevel recombination code. Moreover, some of the key ingredients in the accurate evaluation of the recombination history (e.g. the problem of radiative transfer in hydrogen and the proper inclusion of two-photon processes) are solved using computationally demanding approaches, although in some cases semi-analytical approximations \\citep[see e.g.][]{Hirata2008} might open the possibility of a more efficient evaluation in the future. In order to have an accurate and fast representation of the cosmological recombination history as a function of the cosmological parameters, two possible approaches have been considered in the literature. The first one consists of the inclusion of additional fudge factors to mimic the new physics, as recently done in \\cite{Wong2008} (see {\\sc Recfast} v1.4.2), where they include an additional fudge factor to modify the dynamics of helium recombination. The second approach is the so-called {\\sc Rico} code (FCRW09), which provides an accurate representation of the recombination history by using a regression scheme based on {\\sc Pico} \\citep{Fendt2007a,Fendt2007}. The {\\sc Rico} code smoothly interpolates the $X_{\\rm e}(z; \\vec{p})$ function on a set of pre-computed recombination histories for different cosmologies, where $z$ is the redshift and $\\vec{p}$ represents the set of cosmological parameters. In this paper, we present the results for parameter estimations using {\\sc Rico} with the most recent training set presented in FCRW09. This permits us to accurately account for the full cosmological dependence of the corrections to the recombination history that were included in the multi-level recombination code which was used for the training of {\\sc Rico} (see Sect.~\\ref{sec:rico} for more details). With this tool, we have evaluated the impact of the corrections on the posterior probability distributions that are expected to be obtained for the {\\sc Planck} satellite, by performing a complete Monte Carlo Markov Chain analysis. The study of these posteriors have shown that the impact of the cosmology dependence is not very relevant for those processes included into the current {\\sc Rico} training set. Therefore, by assuming that the cosmology dependence of the correction in general can be neglected, we have also investigated the impact of recent improvements in the treatment of hydrogen recombination (see Sect.~\\ref{sec:Lya}). The basic conclusion is that, with our current understanding of the recombination process, the expected biases in the cosmological parameters inferred from {\\sc Planck} might be as large as 1.5-2.5 sigmas for some parameters as the baryon density or the primordial spectral index of scalar fluctuations, if all these corrections to the recombination history are neglected. The paper is organized as follows. Sect.~\\ref{sec:physics} describes the current training set for {\\sc Rico}, and provides an updated list of physical processes during recombination which were not included in FCRW09. Sect.~\\ref{sec:impact} presents the impact of the recombination uncertainties on cosmological parameter estimation, focusing on the case of {\\sc Planck} satellite. Sect.~\\ref{sec:additional} further extends this study to account for the remaining recombination uncertainties described in Sect.~\\ref{sec:physics}. Sect.~\\ref{sec:current} presents the analysis of present-day CMB experiments, for which the effect is shown to be negligible. Finally, the discussion and conclusions are presented in sections \\ref{sec:discussion} and \\ref{sec:conclusions}, respectively. ", "conclusions": "\\label{sec:conclusions} In this paper, we have performed a MCMC analysis of the expected biases on the cosmological constraints to be derived from the upcoming {\\sc Planck} data, in the light of recent developments in the description of the standard cosmological recombination process. Our main conclusions are: \\begin{itemize} \\item An incomplete description of the cosmological recombination process leads to significant biases (of several sigmas) in some of the basic parameters to be constrained by {\\sc Planck} satellite (see Table~\\ref{tab:planck2}), and in general, by any future CMB experiment. However, these corrections have a minor impact for present-day CMB experiments; for instance, using WMAP5 data plus other cosmological datasets, we find a $\\sim -0.25$ and $\\sim -0.22$ sigma bias on $\\nS$ and $\\logAs$, respectively, while the rest of the parameters remain unchanged. \\item Today, it seems that our understanding of cosmological recombination has reached the sub-percent level in $X_{\\rm e}$ at redshifts $500 \\la z \\la 1600$. However, it will be important to cross-validate all of the considered corrections in a detailed code comparison, which currently is under discussion among the different groups. \\item Given the range of variation of the relevant cosmological parameters, it is possible to incorporate all the new recombination corrections by using (cosmology independent) fudge functions. Here we described one possibility which uses a simple correction factor to the results obtained with {\\sc Recfast} (see Sect.~\\ref{sec:corr_fac}). We provide the function $f(z)$ on the {\\sc Rico}-webpage \\footnote{http://cosmos.astro.uiuc.edu/rico}. \\item The physics of helium recombination already seems to be captured at a sufficient level of precision, when including the acceleration caused by the hydrogen continuum opacity and the $2^3\\rm P_1$-$1^1\\rm S_0$ intercombination line. The biases caused by neglecting only these corrections are -0.8 and -0.4 sigmas, for $\\nS$ and $\\omegab$, respectively. \\item When allowing for more non-standard additions to the recombination model (e.g. related to annihilating dark matter), the biases introduced by an inaccurate recombination model could lead to spurious detections or additional confusion (see Sect.~\\ref{sec:nonstandard}). \\end{itemize}" }, "0910/0910.3330_arXiv.txt": { "abstract": "\\noindent The Carnegie Supernova Project is a five-year survey being carried out at the Las Campanas Observatory to obtain high-quality light curves of $\\sim$100 low-redshift Type~Ia supernovae in a well-defined photometric system. Here we present the first release of photometric data that contains the optical light curves of 35 Type~Ia supernovae, and near-infrared light curves for a subset of 25 events. The data comprise 5559 optical ($ugriBV$) and 1043 near-infrared ($YJHK_s$) data points in the natural system of the Swope telescope. Twenty-eight supernovae have pre-maximum data, and for 15 of these, the observations begin at least 5 days before $B$ maximum. This is one of the most accurate datasets of low-redshift Type~Ia supernovae published to date. When completed, the CSP dataset will constitute a fundamental reference for precise determinations of cosmological parameters, and serve as a rich resource for comparison with models of Type~Ia supernovae. ", "introduction": "\\label{sec:intro} The observation that the expansion rate of the Universe is currently accelerating is arguably one of the most profound discoveries in modern astrophysics. The first direct evidence of this phenomenon was provided a decade ago by the Hubble diagram of high-redshift Type~Ia supernovae \\citep[SNe~Ia;][]{riess98,perlmutter99}. Since then, a wealth of data obtained from surveys of nearby and distant SNe~Ia has confirmed this conclusion, as have other methods such as the X-ray cluster distances \\citep[e.g.,][]{allen07} and the late-time integrated Sachs-Wolfe effect using cosmic microwave background radiation observations \\citep[e.g.,][]{giannantonio08}. These findings suggest that a new form of energy permeates the Universe, or that the theory of General Relativity breaks down on cosmological scales. Today, the major challenge is to determine the nature of this mysterious energy (commonly referred to as ``dark energy'') by measuring its equation-of-state parameter, $w = P/(\\rho c^2)$, and the corresponding time derivative $\\dot w$. SNe~Ia are playing an essential role in the endeavor to measure $w$. Both the SN Legacy Survey \\citep{astier06} and the ESSENCE project \\citep{wood-vasey07} have recently provided independent constraints on $w$ that favor a cosmological constant ($w = -1$). The determination of $\\dot w$, however, will require new and extensive samples of both low- and high-redshift SNe~Ia with systematic errors below 1--2\\%. This will require the construction of a database of low-redshift light curves with excellent photometric precision and temporal coverage The pioneering work of producing such a low-redshift sample of SNe~Ia was carried out by the Cal{\\' a}n/Tololo Survey, which published $BVRI$ light curves of 29 events \\citep{hamuy96}. This is the local sample that was used by \\citet{riess98} and \\citet{perlmutter99} to detect the accelerating expansion of the Universe. In 1999, the CfA Supernova Group released a set of 22 $BVRI$ light curves \\citep{riess99}. Subsequently, this group has published $UBVRI$ light curves of 44 events \\citep{jha06}, and $UBVRIr'i'$ light curves of another 185 \\citep{hicken09}. Other significant samples of nearby SNe~Ia are being produced by the Lick Observatory Supernova Search \\citep[LOSS;][]{filippenko01,filippenko05,filippenko09} and the Nearby Supernova Factory \\citep{aldering02}. In addition, the SDSS-II Supernova Survey \\citep{frieman08} has obtained $u'g'r'i'z'$ light curves of a sample of $\\sim$500 spectroscopically confirmed SNe~Ia in the redshift range of $0.05 < z < 0.35$. In September 2004, the Carnegie Supernova Project (hereafter CSP) began a five-year program to obtain densely sampled optical {\\em and} near-infrared light curves of $\\sim$100 nearby SNe~Ia. The overriding goal has been to obtain the highest possible precision in a stable, well-understood photometric system. An additional objective is to use the broad wavelength leverage afforded by these observations to set stringent constraints on the properties of host-galaxy extinction \\citep[see][]{folatelli09}. In this paper, we present final photometry of the first 35 SNe~Ia observed in the years 2004--2006. This dataset consists of optical ($u'g'r'i'BV$) light curves of all 35 SNe, and near-infrared ($YJHK_s$) light curves for a subset of 25. In total, 5559 optical and 1043 near-infrared measurements were obtained at a typical precision of 0.01--0.03~mag, making this one of the most homogeneous and accurate sets of nearby SN~Ia light curves yet obtained. A distinguishing aspect of the CSP is the quantity and quality of the near-infrared photometry, which is matched only by the recently published PAIRITEL results \\citep{wood08}. Previously, \\citet[][hereafter H06]{hamuy06} gave a detailed description of the CSP observing methodology and data reduction procedures, and showed several examples of light curves obtained during the first campaign. In a second paper, \\citet{folatelli06} reported on the unique Type Ib SN~2005bf, while the third CSP paper presented a detailed analysis of the peculiar Type~Ia SN 2005hk \\citep{phillips07}. A fourth paper on the distance to the Antennae Galaxies (NGC~4038/39) based on the Type~Ia SN~2007sr was presented by \\citet{schweizer08}. CSP photometry has also been included in case studies of the subluminous Type~Ia SN~2005bl \\citep{taubenberger08} and the Type~Ia/IIn SN~2005gj \\citep{prieto07}. Most recently, \\citet{stritzinger09} published a comprehensive study on the Type~Ib SN~2007Y. The organization of this paper is as follows. Section~\\ref{sec:obs} describes the observing and basic data reduction procedures, Section~\\ref{sec:seq} reviews the establishment of the local photometric sequences, Section~\\ref{sec:phot} gives details of the measurement and calibration of the SN photometry, and Section~\\ref{sec:results} presents the final light curves. In two Appendices, we present our most up-to-date information on the sensitivity functions of our $u'g'r'i'BV$ bandpasses, and near-infrared photometry of Feige~16 obtained over the course of our observing campaigns which we use to check the validity of our $Y$-band calibration. Accompanying this paper are two other papers. The first gives an analysis of the data presented here \\citep{folatelli09}. The second combines the data of this article with that of 35 high-redshift SNe~Ia, observed in the near-infrared with the Magellan Baade 6.5-m telescope, in order to construct the first rest-frame $i$-band Hubble diagram out to $z \\approx 0.7$ \\citep{freedman09}. ", "conclusions": "\\label{sec:conc} In this paper, we have presented the first data release of low-redshift SNe~Ia observed by the CSP. The combination of pre-maximum observations, extended coverage, dense sampling, and high signal-to-noise ratio data produces a dataset of unprecedented quality. These results will make possible detailed studies of the photometric properties of SNe~Ia and serve as a rich resource for comparison with theoretical models. When completed, the CSP dataset will constitute a fundamental reference for precise determinations of cosmological parameters such as our own program to measure the $i'$-band Hubble diagram of SNe~Ia to $z \\approx 0.7$ \\citep{freedman09}. The inclusion of the near-infrared data is an important addition since SNe~Ia are nearly perfect standard candles at these wavelengths \\citep{krisciunas04}. In addition, the broad wavelength range covered by the optical and near-infrared observations provides strong leverage for estimating dust extinction in the SN~Ia host galaxies \\citep{folatelli09}." }, "0910/0910.4726_arXiv.txt": { "abstract": "{ In this paper, the stability of a dynamically condensing radiative gas layer is investigated by linear analysis. Our own time-dependent, self-similar solutions describing a dynamical condensing radiative gas layer are used as an unperturbed state. We consider perturbations that are both perpendicular and parallel to the direction of condensation. The transverse wave number of the perturbation is defined by $k$. For $k=0$, it is found that the condensing gas layer is unstable. However, the growth rate is too low to become nonlinear during dynamical condensation. For $k\\ne0$, in general, perturbation equations for constant wave number cannot be reduced to an eigenvalue problem due to the unsteady unperturbed state. Therefore, direct numerical integration of the perturbation equations is performed. For comparison, an eigenvalue problem neglecting the time evolution of the unperturbed state is also solved and both results agree well. The gas layer is unstable for all wave numbers, and the growth rate depends a little on wave number. The behaviour of the perturbation is specified by $kL_\\mathrm{cool}$ at the centre, where the cooling length, $L_\\mathrm{cool}$, represents the length that a sound wave can travel during the cooling time. For $kL_\\mathrm{cool}\\gg1$, the perturbation grows isobarically. For $kL_\\mathrm{cool}\\ll1$, the perturbation grows because each part has a different collapse time without interaction. Since the growth rate is sufficiently high, it is not long before the perturbations become nonlinear during the dynamical condensation. Therefore, according to the linear analysis, the cooling layer is expected to split into fragments with various scales. } ", "introduction": "In the interstellar medium (ISM), it is well known that a clumpy low-temperature phase (cold neutral medium, or CNM) and a diffuse high-temperature phase (warm neutral medium, or WNM) can coexist in pressure equilibrium as a result of the balance of radiative cooling and heating due to external radiation fields and cosmic rays \\citep{FGH69,W95,W03}. These two phases are thermally stable. On the other hand, gas is thermally unstable in the temperature range between two stable phases, that is, in the range 300 K $1$ is also roughly understood from Balbus's criterion. For $1<\\alpha<2$, the gas is isobarically unstable, but it is isochorically stable. For $\\alpha>2$, the gas is thermally stable. In this section, we investigate the stability of the gas for $1<\\alpha<2$ during cooling within the large and small scale limits. \\subsubsection{The isobaric mode} For the case with small wave length, perturbation is expected to grow isobarically. By comparison of our results with previous studies in the literature, it is found that the growth rate in the isobaric mode is independent of the global structure of the unperturbed state. Therefore, from local arguments, the growth rate in the isobaric mode of the gas with $1<\\alpha<2$ can also be estimated. As an unperturbed state, we adopt a cooling gas element whose scale is assumed to be much smaller than the cooling length. In this case, the element cools isobarically. From Eq. (\\ref{eoe}), the time evolution of the unperturbed gas is given by \\begin{equation} \\rho(t)=\\rho_i\\left( 1-\\frac{t}{t_\\mathrm{cool}'} \\right)^{-1/(2-\\alpha)}, \\frac{1}{t_\\mathrm{cool}'} = (2-\\alpha)\\gamma^{\\alpha-1}(\\gamma-1) P_i^{\\alpha-1}\\rho_i^{2-\\alpha}, \\label{a ge 1} \\end{equation} where $\\rho_i$ and $P_i$ represent the initial density and pressure, respectively. In the above unperturbed state, we consider the following isobaric perturbation: \\begin{equation} \\rho = \\rho_0(t) +\\delta \\rho(t), \\end{equation} and \\begin{equation} P = P_0, \\end{equation} where subscript ``0'' indicates the unperturbed state, and $\\delta \\rho$ is the density perturbation. Linearizing Eq. (\\ref{eoe}), one obtains \\begin{equation} \\frac{\\mathrm{d} }{\\mathrm{d}t}\\left( \\frac{\\delta\\rho}{\\rho_0} \\right) = (2-\\alpha)\\gamma\\gamma^{\\alpha-1} \\rho_0^{2-\\alpha}P_0^{\\alpha-1}\\frac{\\delta \\rho}{\\rho_0}. \\label{per a ge 1} \\end{equation} Using Eq. (\\ref{a ge 1}), Eq. (\\ref{per a ge 1}) is rewritten as \\begin{equation} \\frac{\\mathrm{d} }{\\mathrm{d}t}\\left( \\frac{\\delta\\rho}{\\rho_0} \\right) = \\frac{1}{t_\\mathrm{cool}'} \\left( 1-\\frac{t}{t_\\mathrm{cool}'} \\right)^{-1} \\frac{\\delta \\rho}{\\rho_0}. \\label{per a ge 1 2} \\end{equation} Equation (\\ref{per a ge 1 2}) is easily integrated to give \\begin{equation} \\frac{\\delta \\rho}{\\rho_0} \\propto \\left( 1-\\frac{t}{t_\\mathrm{cool}'} \\right)^{-1}. \\label{a ge 1 growth rate} \\end{equation} Comparing Eq. (\\ref{a ge 1 growth rate}) with Eq. (\\ref{a ge 1}), one can see that the perturbation grows more slowly than the unperturbed state for $1<\\alpha<2$. Therefore, the gas is expected to be difficult to fragment during runaway cooling if $1<\\alpha<2$. \\subsubsection{The noninteractive mode} A cooling layer that evolves isobarically is considered. The time evolution of the central density is the same as Eq. (\\ref{a ge 1}). When the scale of perturbation perpendicular to the condensation is too large to interact with other regions, each region evolves independently. Here, we focus on the time evolution of density perturbation at the centre ($x=0$). Initial fluctuation of the central density, $\\delta \\rho_i$, creates the fluctuation of the cooling time, $\\Delta t$. The relative amplitude of density perturbation at $x=0$ is given by \\begin{equation} \\frac{\\delta \\rho}{\\rho_0} = \\frac{1}{\\rho_0}\\left( \\rho_i + \\delta \\rho_i \\right) \\left( 1-\\frac{t}{t_\\mathrm{cool}' - \\Delta t} \\right)^{-1/(2-\\alpha)}-1. \\label{non int a ge 1} \\end{equation} Linearizing Eq. (\\ref{non int a ge 1}) with omitting terms that do not grow, we have \\begin{equation} \\frac{\\delta \\rho}{\\rho_0} = \\frac{1}{2-\\alpha}\\frac{\\Delta t}{t_\\mathrm{cool}'} \\left( 1-\\frac{t}{t_\\mathrm{cool}'} \\right)^{-1}. \\label{non in a ge 1 1} \\end{equation} Comparing Eq. (\\ref{non in a ge 1 1}) with Eq. (\\ref{a ge 1}), we can see that the perturbation grows more slowly than the unperturbed state for $1<\\alpha<2$. Therefore, the gas is expected to be difficult to fragment for $1<\\alpha<2$ for the large-scale perturbation, as well as the small scale. \\subsection{Effects of thermal conduction}\\label{thermal conduction} In this paper, the thermal conduction is neglected for simplicity. However, for large wave number, the thermal conduction is expected to stabilize TI in the cooling layer \\citep{F65}. Therefore, there is a critical wave number, $k_\\mathrm{crit}$, such that perturbation with a larger wave number is stabilized by the thermal conduction. First, we evaluate $k_\\mathrm{crit}$ using an order estimation. Using the characteristic time scale of the thermal conduction, $t_\\mathrm{diff}$, the diffusion equation is given by \\begin{equation} \\frac{1}{t_\\mathrm{diff}} \\left( \\frac{P_{00}}{\\gamma-1} \\right) \\sim k^2K(T_{00}) T_{00}, \\label{diffusion eq} \\end{equation} where $K$ is the thermal conduction coefficient. From Eq. (\\ref{diffusion eq}), the diffusion timescale is given by \\begin{equation} t_\\mathrm{diff} \\simeq \\frac{P_{00}}{(\\gamma-1)K(T_{00})T_{00} } k^{-2}. \\label{diffusion} \\end{equation} From Eq. (\\ref{diffusion}), one can see that the diffusion timescale is small for large wave number. If $t_\\mathrm{diff}k_\\mathrm{crit}$) becomes unstable at a certain epoch when $k_\\mathrm{cirt}$ catches up with $k$. Figure \\ref{alpha evo} shows the time evolution of the critical wave length, $\\lambda_\\mathrm{crit}=2\\pi/k_\\mathrm{crit}$. In Sect. \\ref{astro impli}, the time evolution of the condensing layer can be described by the S-S solution with ($\\alpha_\\mathrm{num}=0.61$, $\\eta=0.98$). In Fig. \\ref{alpha evo}, we see that $\\lambda_\\mathrm{crit}$ decreases with time during the runaway condensation. Figure \\ref{alpha evo} also shows that $\\lambda_\\mathrm{crit}1$. This means that the effect of thermal conduction on the dispersion relation (Fig. \\ref{kneq0 grow}) always appears in the isobaric regime during the runaway condensation." }, "0910/0910.0333_arXiv.txt": { "abstract": "{ Over almost all of minimal supergravity (mSUGRA or CMSSM) model parameter space, there is a large overabundance of neutralino cold dark matter (CDM). We find that the allowed regions of mSUGRA parameter space which match the measured abundance of CDM in the universe are highly fine-tuned. If instead we invoke the Peccei-Quinn-Weinberg-Wilczek solution to the strong $CP$ problem, then the SUSY CDM may consist of an axion/axino admixture with an axino mass of order the MeV scale, and where mixed axion/axino or mainly axion CDM seems preferred. In this case, fine-tuning of the relic density is typically much lower, showing that axion/axino CDM ($a\\tilde{a}$CDM) is to be preferred in the paradigm model for SUSY phenomenology. For mSUGRA with $a\\tilde{a}$CDM, quite different regions of parameter space are now DM-favored as compared to the case of neutralino DM. Thus, rather different SUSY signatures are expected at the LHC in the case of mSUGRA with $a\\tilde{a}$CDM, as compared to mSUGRA with neutralino CDM. } ", "introduction": "\\label{sec:intro} A wide array of astrophysical data point to us living in a universe comprised of $4\\%$ baryons, $\\sim 25\\%$ cold dark matter (CDM) and $\\sim 70\\%$ dark energy. In fact, the cosmic abundance of CDM has been recently measured to high precision by the WMAP collaboration\\cite{wmap5}, which finds \\be \\Omega_{CDM}h^2=0.110\\pm 0.006 , \\ee where $\\Omega=\\rho/\\rho_c$ is the dark matter density relative to the closure density, and $h$ is the scaled Hubble constant. No particle present in the Standard Model (SM) of particle physics has the correct properties to constitue the CDM, so some form of new physics is needed. It is compelling, however, that candidate CDM particles do emerge naturally from two theories which provide solutions to longstanding problems in particle physics. The first problem-- known as the gauge hierarchy problem-- arises due to quadratic divergences in the scalar sector of the SM. These divergences lead to scalar masses blowing up to the highest scale in the theory ({\\it e.g.} in grand unified theories (GUTS), the GUT scale $M_{GUT}\\simeq 2\\times 10^{16}$ GeV), unless an enormous fine-tuning of parameters is invoked. One solution to the gauge hierarchy problem occurs by introducing supersymmetry (SUSY) into the theory. The inclusion of softly broken SUSY leads to a cancellation of quadratic divergences between fermion and boson loops, so that only log divergences remain. The log divergence is soft enough that vastly different scales remain stable within a single effective theory. In SUSY theories, the lightest neutralino emerges as an excellent WIMP CDM candidate. Gravity-mediated SUSY breaking models (supergravity, or SUGRA) contain gravitinos with weak-scale masses. SUGRA models experience tension due to possible overproduction of gravitinos in the early universe, leading to an overabundance of CDM. In addition, gravitinos usually decay during or after Big Bang nucleosynthesis (BBN), and their energetic decay products may disrupt the successful calculations of light element abundances, which otherwise maintain good agreement with observation. This tension in SUGRA models is known as the {\\it gravitino problem}. The second problem is the strong $CP$ problem\\cite{kcreview}. An elegant solution to the strong $CP$ problem was proposed by Peccei and Quinn (PQ) many years ago\\cite{pq}. The PQ solution automatically predicts the existence of a new particle (WW)\\cite{ww}: the axion $a$. While the original PQWW axion was soon ruled out, models of a nearly ``invisible axion'' were developed in which the PQ symmetry breaking scale was moved up to energies of order $f_a\\sim 10^{9}-10^{12}$ GeV\\cite{ksvz,dfsz}. The axion also turns out to be an excellent candidate particle for CDM in the universe\\cite{absik}. Of course, it is highly desirable to simultaneously account for {\\it both} the strong $CP$ problem and the gauge hierarchy problem. In this case, it is useful to invoke supersymmetric models which include the PQWW solution to the strong $CP$ problem\\cite{nillesraby}. In a SUSY context, the axion field is just one element of an {\\it axion supermultiplet}. The axion supermultiplet contains a complex scalar field, whose real part is the $R$-parity even saxion field $s(x)$, and whose imaginary part is the axion field $a(x)$. The supermultiplet also contains an $R$-parity odd spin-$\\frac{1}{2}$ Majorana field, the axino $\\ta$\\cite{steffen_rev}. The saxion, while being an $R$-parity even field, nonethless receives a SUSY breaking mass likely of order the weak scale. The axion mass is constrained by cosmology and astrophysics to lie in a favored range $10^{-2}$ eV$\\agt m_a\\agt 10^{-5}$ eV. The axino mass is very model dependent\\cite{axmass,rtw,cl,ckkr}, depending heavily on the exact form of the superpotential and the mechanism for SUSY breaking. In supergravity models, it may be of order the gravitino mass $m_{3/2}\\sim$ TeV, or as low as $m_{3/2}^2/f_a\\sim $keV. Conditions for realizing these extremes are addressed in \\cite{cl}. Here, we will try to avoid explicit model-dependence, and adopt $m_{\\ta}$ as lying within the general range of keV-GeV, as in numerous previous works\\cite{rtw,ckkr,fstw,cmssm,axdm}. An axino in this mass range would likely serve as the lightest SUSY particle (LSP), and is also a good candidate particle for cold dark matter\\cite{rtw,ckkr}. In a previous paper\\cite{axdm}, we investigated supersymmetric models wherein the PQ solution to the strong $CP$ problem is assumed. For definiteness, we restricted the analysis to examining the paradigm minimal supergravity (mSUGRA or CMSSM) model\\cite{msugra}. We were guided in our analysis by considering the possibility of including a viable mechanism for baryogenesis in the early universe. In order to do so, we needed to allow for re-heat temperatures after the inflationary epoch to reach values $T_R\\agt 10^6$ GeV. We found that in order to sustain such high re-heat temperatures, as well as generating predominantly {\\it cold} dark matter, we were pushed into mSUGRA parameter space regions that are very different from those allowed by the case of thermally produced neutralino dark matter. In addition, we found that very high values of the PQ breaking scale $f_a/N$ of order $10^{11}-10^{12}$ GeV were needed, leading to the mSUGRA model with {\\it mainly axion cold dark matter}, but also with a small admixture of thermally produced axinos, and an even smaller component of warm axino dark matter arising from neutralino decays. The favored axino mass value is of order 100 keV. We note here recent work on models with dominant axion CDM explore the possibility that axions form a cosmic Bose-Einstein condensate, which can allow for the solution of several problems associated with large scale structure and the cosmic background radiation\\cite{pierre}. In this paper, we will examine the mSUGRA model under the assumption 1. of neutralino CDM and 2. that mixed axion/axino DM ($a\\tilde{a}$DM) saturates the WMAP measured abundance\\footnote{The possibility of mixed $a\\tilde{a}$CDM was suggested in the context of Yukawa-unified SUSY in Ref. \\cite{mix}.}. To compare the two DM scenarios, we will evaluate a measure of fine-tuning in the relic abundance \\be \\Delta_{a_i} \\equiv\\frac{\\partial\\log\\Omega_{DM}h^2}{\\partial\\log a_i} \\ee with respect to variations in fundamental parameters $a_i$ of the model. Such a measure of relic abundance fine-tuning was previously calculated in Ref.~\\cite{eo} in the context of just neutralino dark matter. Here, we will expand upon this and also consider fine-tuning of the relic density in the case of mixed $a\\tilde{a}$DM. Our main conclusion is that the relic abundance of DM is {\\it much less fine-tuned in the case of mixed $a\\tilde{a}$CDM, as compared to neutralino CDM}. Thus, we find that mixed $a\\tilde{a}$CDM is {\\it theoretically preferable to neutralino CDM}, at least in the case of the mSUGRA model, and probably also in many cases of SUGRA models with non-universal soft SUSY breaking terms. We will restrict our work to cases where the lightest neutralino $\\tz_1$ is either the LSP or the next-to-lightest SUSY particle (NLSP) with an axino LSP; the case with a stau NLSP and an axino LSP has recently been examined in Ref. \\cite{fstw}. Related previous work on axino DM in mSUGRA can be found in Ref. \\cite{cmssm}. The remainder of this paper is organized as follows. In Sec. \\ref{sec:inoDM}, we calculate the neutralino relic abundance fine-tuning parameter $\\Delta_{\\tz_1}$ in the mSUGRA model due to variation in parameters $m_0$ and $m_{1/2}$. We find, in good agreement with Ref. \\cite{eo}, that the WMAP allowed regions are all finely-tuned for low values of $\\tan\\beta$. For much higher $\\tan\\beta\\sim 50$, the fine-tuning is much less with respect to $m_0$ and $m_{1/2}$, but nevertheless high with respect to $\\tan\\beta$. In Sec. \\ref{sec:inoprob}, we review the gravitino problem, leptogenesis and the cosmological production of axion and axino dark matter. In Sec. \\ref{sec:axDM}, we calculate the fine-tuning parameter $\\Delta_{a\\ta}$ for mixed $a\\tilde{a}$CDM under the assumption of a very light axino with $m_{\\ta}\\sim 0.1-1$ MeV. The fine-tuning is always quite low, for both cases of mixed axino/axion CDM and mainly axion CDM. In the case of mainly axino CDM, we find the scenario less well-motivated since for high values of $T_R\\agt 10^6$ GeV, the value of $m_{\\ta}\\ll 0.1$ MeV, making the axino mainly {\\it warm} DM instead of cold DM. In Sec. \\ref{sec:conclude}, we present a summary and conclusions. ", "conclusions": "\\label{sec:conclude} In this paper, we have examined the fine-tuning associated with the relic density of dark matter in the minimal supergravity model. We have calculated a measure of fine-tuning assuming two scenarios for SUSY dark matter: 1. neutralino dark matter with fine-tuning parameter $\\Delta_{\\tz_1}$, and 2. mixed axion/axino dark matter with fine-tuning parameter $\\Delta_{a\\ta}$. In the case of neutralino dark matter, we find that the WMAP-allowed regions of mSUGRA such as the stau co-annihilation region, the HB/FP region and the light Higgs $h$-resonance annihilation region, are all rather highly fine-tuned, especially the HB/FP region, where $\\Delta_{\\tz_1}$ ranges from 20-100. Only mild fine-tuning is found in the low $m_0$, low $m_{1/2}$ region where stau co-annihilation and bulk annihilation through $t$-channel slepton exchange overlap. If one moves to large $\\tan\\beta\\sim 50$, then larger regions of parameter space which are consistent with WMAP occur. These large $\\tan\\beta$ regions have modest fine-tuning versus $m_0$ and $m_{1/2}$, but very large fine-tuning versus $\\tan\\beta$. If instead we assume that dark matter is composed of an axion/axino admixture, rather than neutralinos, then we find that the relic density fine-tuning parameter is generically much lower: $\\Delta_{a\\ta}\\sim 1.3-2.5$ throughout parameter space. Here, we have assumed the existence of a light axino with mass $m_{\\ta}\\sim$ keV-MeV. Such a light axino opens up all of mSUGRA parameter space to being WMAP allowed, since now the neutralino decays via $\\tz_1\\to \\ta\\gamma$. If the DM is dominated by thermally produced axinos, then the re-heat temperature $T_R$ is generally lower than $10^6$ GeV unless the axinos are actually warm dark matter ($m_{\\ta}\\alt 100$ keV), so this scenario seems rather unlikely. However, if the PQ breaking scale $f_a/N$ is large, then the DM can be either a nearly equal axion/axino admixture, in which case fine-tuning is lowest ($\\Delta_{a\\ta}\\sim 1.3$), or a dominantly axion mixture (in which case $\\Delta_{a\\ta}\\sim 2.3$). Either scenario easily admits $T_R>10^6$ GeV, which can allow for non-thermal leptogenesis to occur. The consequences of the mixed $a\\tilde{a}$CDM scenario for future dark matter searches is as follows. For collider searches, we expect much the same collider signatures as in the mSUGRA model with neutralino dark matter, since we assume the $\\tz_1$ is the NLSP, and decays far outside the collider detectors. However, {\\it all} of mSUGRA parameter space is now WMAP-allowed, instead of just the special co-annihilation, HB/FP region and resonance annihilation regions. As shown in Ref. \\cite{axdm}, the regions of WMAP-allowed neutralino CDM yield the lowest values of $T_R$, and so the stau co-annihilation, HB/FP region and $h$ resonance annihilation regions are most dis-favored for the case of mixed $a\\ta$CDM. As far as WIMP searches go, in the mixed $a\\tilde{a}$CDM scenario, we expect no positive signals if $m_{\\tz_1}>m_{\\ta}$. If $m_{\\ta}>m_{\\tz_1}$, then the $\\tz_1$ would still be stable (assuming $R$-parity conservation) and WIMP direct and indirect detection signals are still possible\\cite{njp}. In the case of large axion relic abundance, which appears to us to be the favored scenario, then a positive signal at relic axion search experiments such as ADMX might be expected\\cite{admx}, although solar axion searches are less likely to achieve positive results, since large values of $f_a/N$ are favored, leading to small axion/axino couplings. Our analysis has been based on the admittedly subjective basis of fine-tuning of the relic density of dark matter relative to model input parameters. We note here that the mSUGRA model already needs substantial fine-tuning in the electroweak sector in order to accomodate the relatively light $Z$ boson mass in the face of limits on the soft SUSY breaking parameters\\cite{ewft} (the little hierarchy problem). Our philosophy here is that less fine-tuning is better, and high fine-tuning in one sector is better than high fine-tuning in two sectors, {\\it e.g.} electroweak and dark matter sectors. While our analysis has been restricted to the mSUGRA SUSY model, one might ask how general our conclusions might be. In SUSY models based on gravity mediation, with a neutralino LSP, the DM relic density is {\\it generically too high} unless some special mechanism is acting to enhance the neutralino annihilation cross section in the early universe.\\footnote{ Discussion on numerous different SUGRA models with non-universality has been explored in Ref. \\cite{wtn}.} For instance, in SUSY models with non-universality, instead of stau or stop co-annihilation, one might have sbottom or sneutrino co-annihilation, or bino-wino co-annihilation: in any case, the mass gap between co-annihilating particles must be fine-tuned to obtain agreement with the measured dark matter abundance. In non-universal models with a {\\it well-tempered neutralino}\\cite{wtn}, where the neutralino bino-higgsino or bino-wino composition is adjusted to fit the measured relic density, other parameters (Higgs soft masses, gaugino masses) must be fine-tuned to get just the right ``tempering'', as occurs in the mSUGRA HB/FP region. In other models, Higgs soft mass terms can be adjusted to allow $2m_{\\tz_1}$ to sit atop the $A$ resonance; but again, in this case, parameters must be fine-tuned (unless $\\tan\\beta$ is large, which also occurs in mSUGRA). The case where the SUSY neutralino abundance is not fine-tuned has long been noted: it is where squarks and sleptons are so light that $t$-channel annihilation channels are large. However, LEP2 search limits now essentially exclude all these regions. Thus, although we restrict our analysis here to the mSUGRA model, we feel this model provides a sort of microcosm for general SUSY models, in that it illustrates many of the features common to all SUSY models. Our main conclusion is this. In the world HEP community, a tremendous effort is underway to explore for WIMP cold dark matter, based partly on the view that SUSY models naturally give rise to the ``WIMP-miracle'', and an excellent WIMP candidate for CDM. We have shown here that at least for the paradigm SUSY model-- mSUGRA-- usually a large overabundance of neutralino CDM is produced, unless one lies along a region of very high fine-tuning, where a slight change in model parameters leads to a large change in relic density: this equates to a high degree of relic density fine-tuning. Alternatively, if one assumes the PQWW solution to the strong CP problem within SUSY models, and a very light axino with $m_{\\ta}$ of the order of MeV, then along with an elegant solution to the strong CP problem, one obtains a mixed axion/axino relic density with much less fine-tuning. Given our results, we would advocate that a much increased share of HEP resources be given to relic axion searches, where the global search effort has been much more limited." }, "0910/0910.5720_arXiv.txt": { "abstract": "We present a new analysis of stellar mass functions in the COSMOS field to fainter limits than has been previously probed at $z \\leq 1$. The increase in dynamic range reveals features in the shape of the stellar mass function that deviate from a single Schechter function. Neither the total nor the red (passive) or blue (star-forming) galaxy stellar mass functions can be well fit with a single Schechter function once the mass completeness limit of the sample probes below $\\sim 3 \\times 10^{9} \\Msun$. We observe a dip or plateau at masses $\\sim 10^{10} \\Msun$, just below the traditional $M^*$, and an upturn towards a steep faint-end slope of $\\alpha \\sim -1.7$ at lower mass at all redshifts $\\leq 1$. This bimodal nature of the mass function is {\\em not} solely a result of the blue/red dichotomy. Indeed, the blue mass function is by itself bimodal at $z \\sim 1$. This suggests a new dichotomy in galaxy formation that predates the appearance of the red sequence. We propose two interpretations for this bimodal distribution. If the gas fraction increases towards lower mass, galaxies with $M_{\\rm baryon} \\sim 10^{10} \\Msun$ would shift to lower stellar masses, creating the observed dip. This would indicate a change in star formation efficiency, perhaps linked to supernovae feedback becoming much more efficient below $\\sim$10$^{10} \\Msun$. Therefore, we investigate whether the dip is present in the baryonic (stars+gas) mass function. Alternatively, the dip could be created by an enhancement of the galaxy assembly rate at $\\sim$10$^{11} \\Msun$, a phenomenon that naturally arises if the baryon fraction peaks at $M_{\\rm halo} \\sim 10^{12} \\Msun$. In this scenario, galaxies occupying the bump around $\\Mstar$ would be identified with central galaxies and the second fainter component of the mass function having a steep faint-end slope with satellite galaxies. The low-mass end of the blue and total mass functions exhibit a steeper slope than has been detected in previous work that may increasingly approach the halo mass function value of -2. While the dip feature is apparent in the total mass function at all redshifts, it appears to shift from the blue to red population, likely as a result of transforming high-mass blue galaxies into red ones. At the same time, we detect a drastic upturn in the number of low-mass red galaxies. Their increase with time seems to reflect a decrease in the number of blue systems and so we tentatively associate them with satellite dwarf (spheroidal) galaxies that have undergone quenching due to environmental processes. ", "introduction": "\\label{sec:introduction} Galaxy formation and evolution is believed to be driven primarily by two processes: firstly, the successive merging of their parent dark matter halos causing accretion of material and ultimately mergers between galaxies; and secondly, the feedback-regulated conversion of gas into stars within galactic disks with subsequent potential rearrangement of the disk material by dynamical processes (secular evolution). Both processes contribute to the growth in stellar mass of galaxies with time. The stellar mass function of galaxies and its evolution with time is therefore fundamental to the understanding of galaxy formation. The ability to estimate galaxy stellar masses has advanced in recent years in large part because of increasing access to near-IR photometry. Estimates are typically made by fitting multi-band photometry with stellar population synthesis libraries (see, e.g., \\citealp{Brinchmann+2000,Bruzual+2003,Drory+2003,Drory+2004,Maraston+2006,Marchesini+2008,Conroy+2009}), fitting specific spectral features when spectroscopy is available \\citep{Kauffmann+2003}, or the full spectrum when high-quality spectra are observed \\citep{Reichardt+2001,Panter+2004}. So far, only the photometric fitting technique has been a viable option for high-redshift surveys. These measurements provide masses with accuracies of $\\sim 0.1-0.3$~dex, and systematic uncertainties of up to a factor of two depending on the selection and number of photometric bands included and the assumptions made on, among others, the shape of the IMF, the allowed star formation histories, the dust extinction model, or the underlying stellar population synthesis method (see, e.g., \\citealp{Drory+2004,Kannappan+2007,Marchesini+2008,Conroy+2009} for systematic studies on this matter). Utilizing such techniques, the build-up of the stellar mass density from redshift $z \\sim 6$ to the present epoch has been the subject of several studies in the past decade, often relying on deep multi-band imaging surveys in the UV to near-infrared wavelength range \\citep{Brinchmann+2000,Drory+2001a,Cohen2002,Dickinson+2003,Fontana+2003,Rudnick+2003,Glazebrook+2004,Conselice+2005,Chapman+2005,Drory+2005a,Rudnick+2006,Eyles+2007,Grazian+2007,Stark+2007}. The stellar mass function in the local universe has been measured from large imaging and spectroscopic surveys such as 2dF, SDSS, and 2MASS \\citep{Cole+2001,Bell+2003,Perez-Gonzalez+2003,Panter+2004,Baldry+2008}. At distances up to $z \\sim 1.5$, a number of groups have established a picture of the evolution of the mass function with some detail \\citep{Drory+2003,Fontana+2004,Bundy+2005,Borch+2006,Bundy+2006,Arnouts+2007,Pozzetti+2007,Ilbert+2009}, with generally good agreement between different datasets. To some lesser detail and accuracy, deep surveys have provided data spanning $0 < z \\lesssim 5$ \\citep{Drory+2005a,Conselice+2005,Fontana+2006,Yan+2006,Grazian+2007,Elsner+2008,Perez-Gonzalez+2008,Marchesini+2008}, and even some estimates at $z \\sim 7$ \\citep{Bouwens+2006}. So far, these high-redshift studies of the stellar mass function emphasized the evolution of galaxies of mass $\\logMsun \\gtrsim 10$. Speaking very broadly, the stellar mass density decreases by a factor of two to $z \\sim 1$, with the most massive galaxies already being in place at earlier epochs. The evolution appears to accelerate quickly beyond $z \\sim 1.5$. In this paper, we concentrate on the low-mass galaxies that have typically been below the completeness limits of previous work at $z > 0.1$. Generally \\citet{Schechter1976} fits to the galaxy stellar mass function with faint-end slope $\\sim -1.1$ to $\\sim -1.3$ have been found adequate to describe the galaxy population (even separated morphologically, by color, or star formation activity; \\citealp{Pannella+2006,Borch+2006,Arnouts+2007,Pannella+2009,Ilbert+2009}). Recently, though, a steepening of the slope of the luminosity function below $M_i \\sim -17$ in the local universe has been convincingly detected in clusters \\citep{Driver+1994a,Trentham+2002,Hilker+2003,Popesso+2005,Popesso+2006}, groups \\citep{Trentham+2002,Trentham+2005,Gonzalez+2006}, and in the field \\citep{Blanton+2005a}. For example, \\citet{Trentham+2002} find that the luminosity function in the Virgo cluster, in the NGC~1407 group, and in the Coma~1 group is steep between $M_R$ of -18 and -15 (and flattens again only at $M_R > -15$). Moreover, \\citet{Baldry+2008} find that the local galaxy stellar mass function steepens as well below $\\logMsun \\sim 9.5$ (but see also \\citealp{Li+2009}). The steepening of the mass function can also be interpreted as a bimodality: the mass function consists of a sum of (at least) two components. This bimodal behavior has now also been detected at redshifts $z>0.1$. \\citet{Pozzetti+2009} find bimodal mass functions to $z\\sim 0.5$ from the zCOSMOS spectroscopic survey. They interpret the mass function as being composed of early-type galaxies dominating the massive part and late-type galaxies dominating the less massive part and contributing the steep faint-end slope. Each of these components is well fit by a Schechter function. \\citet{Bolzonella+2009} use the same sample to investigate the bimodality as a function of environment. They find that at $z\\lesssim 0.5$, the shape of the galaxy stellar mass function in high- and low-density environments become markedly different, with high density regions showing a stronger bimodality. We extend the study of the shape of the galaxy stellar mass function, particularly at low masses, to $z \\sim 1$, with stellar mass limits $\\sim 1.5$~dex lower than can be achieved with spectroscopic studies. We investigate actively star-forming galaxies and passively evolving galaxies separately; we study how the change in slope may be caused by the presence of multiple galaxy populations, that taken together, lead to a mass function shape that is more complex than a single power law with an exponential cutoff or even a simple combination of early- and late-type components. We show that the blue mass function itself is bimodal and that passive galaxies exhibit a faint-end upturn, likely caused by dwarf spheroidal galaxies linked to the faint end of the blue galaxy population. This paper is organized as follows: in \\S~\\ref{sec:sample} we introduce the galaxy sample that we use in this work. In \\S~\\ref{sec:stellar-mass} we discuss the stellar population models used to derive stellar masses and the resulting mass completeness limits. In \\S~\\ref{sec:mf} we present the stellar mass function of active and passive galaxies and we discuss our results in \\S~\\ref{sec:discuss}. Finally, we summarize this work in \\S~\\ref{sec:summary}. Throughout this work we assume $\\Omega_M = 0.3$, $\\Omega_{\\Lambda} = 0.7, H_0 = 70\\ h_{70}^{-1} \\mathrm{km\\ s^{-1}\\ Mpc^{-1}}$. Magnitudes are in the AB system. We will denote galaxy stellar masses by the symbol $M$ -- or $\\Mg$ where an explicit distinction form halo masses, denoted by $\\Mh$, is necessary. The symbol $M^*$ is reserved for the characteristic mass parameter of the Schechter function. ", "conclusions": "\\label{sec:discuss} \\subsection{Comparison to Previous Results} \\begin{figure*}[ht] \\centering \\includegraphics[width=0.8\\textwidth]{fig8.eps} \\caption{A comparison of the mass function deduced in this work with data from the literature. The mass function of passive galaxies is marked in red, that of star-forming galaxies in blue, and that of all galaxies in black. The solid lines show double Schechter function fits to the data (Eq.~\\ref{eq:schechter2}; the thin lines show the individual bright and faint components of the double Schechter function). The fit at low $z$ is repeated in higher-$z$ panels as dotted lines. The magenta line denotes the \\citet{Baldry+2008} $z \\sim 0$ SDSS-based mass function convolved with the photo-z error distributions (see text). The dashed blue, red, and black lines are the 3.6$\\mu$m-selected mass functions of active, passive, and all galaxies, respectively, in the COSMOS survey by \\citet{Ilbert+2009}. The green dash-dotted line is the mass function by \\citet{Perez-Gonzalez+2008}. Additionally, we plot the Press-Schechter halo mass function scaled by the global baryon fraction $f_{\\rm b}$ (\\citealp{Dunkley+2009}; WMAP5) in cyan. \\label{fig:mf_lit}} \\end{figure*} Before discussing possible interpretations of the more complicated mass function shape apparent in the COSMOS data, it is useful to compare our results to previous work with the aim of determining whether evidence for similar behavior has been found in other surveys. In Fig.~\\ref{fig:mf_lit} we compare our mass function to a number of results from the literature and for later reference we include the halo mass function \\citep{Reed+2007} where the halo masses have been multiplied by the global baryon fraction, $f_{\\rm b} = \\Omega_{\\rm b}/\\Omega_{\\rm m}$, taken from the WMAP 5-yr data \\citep{Dunkley+2009}. We show the $z \\sim 0$ SDSS-based mass function by \\citep{Baldry+2008} convolved with the photo-z error distributions as shown in Fig.~\\ref{fig:mc_conv} and discussed in \\S~\\ref{sec:photoz_uncertainties}. We also plot the mass functions in the redshift range $0 1$ many more are likely to be rapidly forming stars), the dip in the stellar mass function should move to higher masses at early times as well. Unfortunately, the mass functions presented here are likely to be too affected by cosmic variance to robustly measure redshift evolution in the dip feature, but future data sets will be able to overcome this problem. Also, a more rigorous analysis taking into account cosmological gas accretion as well as a realistic feedback prescription and outflow model is needed, but is beyond the scope of this paper. {\\bf Hierarchical assembly.} Another possibility is that the dip at $\\log M \\sim 10$ and the bump around $M^*$ are formed by a depletion of stellar and gas mass at intermediate mass resulting from a pile-up of the end-products of mass-dependent assembly around $M^*$. In this scenario, both the stellar and baryonic mass functions would exhibit dips at $\\log M \\sim 10$ because the assembly of baryonic mass is accelerated at a certain scale. It is known that the halo-halo merger rate depends very weakly on mass \\citep{Fakhouri+2008}. However, if the stellar mass, or in this case, baryonic mass fraction, peaks in halos at a given mass scale---perhaps as a result of the SFR feedback discussed above---the resulting merger rate in terms of stellar mass, $M$, or baryonic mass, $M_{\\rm baryons}$, can be enhanced near this scale \\citep{Bundy+2009,Stewart2009,Hopkins+2009a} and strengthen the dip feature. This is because the dynamical friction timescale (and hence the merger timescale) depends firstly on the mass ratio of the merging systems (e.g.\\ \\citealp{Boylan-Kolchin+2008} and references therein). Low mass ratio mergers happen only after a long time, and one-to-one mergers happen quickly. At low mass, one-to-one mergers in DM correspond to minor mergers in baryons and hence the baryonic content of a halo assembles more slowly than the halo. This is due to the steep relation between stellar mass, \\Mg, and halo mass, \\Mh, below $M^*$ (see Fig.~\\ref{fig:abmatch}). At masses near and just above the scale at which $\\Mg(\\Mh)$ flattens, more frequent minor halo mergers can host one-to-one baryonic mergers and thus the baryonic assembly rates increase markedly. At even higher mass scales, above the flattening at $M^*$ of the $\\Mh(\\Mg)$-relation, the infalling galaxy is much more likely to become and stay a satellite, even if the baryonic mass ratio is close to one. At this point, the baryonic merger rate drops again. As a result, low stellar (baryonic) mass galaxies get depleted more rapidly in one-to-one mergers, and pile-up around the $M^*$ mass scale ($M^* \\sim 10^11~\\Msun$). This scenario also suggests that the two populations of galaxies might be thought of as central galaxies occupying the bump with an increasing fraction of satellite galaxies at lower masses that possibly form the second, faint, component of the mass function, leading to a steepening of the low-mass end of the mass function. Such satellites in a galaxy or group halo would orbit for a long time before merging and hence remain visible as individual galaxies. If this enhanced assembly scenario were important, it would also indicate that many major mergers between star-forming (disk) galaxies must not lead to a final destruction of the disk and truncation of star formation as the bump signature is already apparent in the blue galaxy population at $z\\sim 1$. Such mergers must produce a remnant that is still a blue star-forming disk galaxy. Such outcomes of mergers seem possible if the gas fraction is large \\citep{Springel+2005a,Robertson+2006,Governato+2007,Hopkins+2009b}. Also, we would expect the dip feature to deepen with time but, in contrast with SFR scenario above, remain relatively fixed with respect to mass. {\\bf Steepening Faint-End Slope.} So far we have focused on the dip in the stellar mass function. We now turn to the steep slopes at the low-mass end. To the extent that this feature is unrelated to the dip forming process mentioned above, we can consider several possibilities. To begin with, many authors interpret the faint-end slope as rising to the maximum set by the halo mass function (with a value of -2). Such behavior is expected if feedback processes that regulate the star formation efficiency become increasingly independent of $M_{\\rm halo}$ below some scale. At low masses, for example, supernova feedback may so dominate over the ability of halos to retain gas, that the effective gas or stellar mass fraction becomes a constant, independent of halo mass. \\subsubsection{Red Galaxies}\\label{sec:red} We now turn to the behavior of the red galaxy mass function. In the redshift bins $z=0.5$ and $z=0.3$, where the faint red component of the mass function is sampled well, the number density of all galaxies at $M \\lesssim 9.5$ does not change much (Fig.~\\ref{fig:mf_lit} and Table~\\ref{tab:mffit}). Yet, the number density of faint red galaxies increases by 0.45~dex, a combined effect of the normalization increasing from $\\phif(z=0.5) = 0.28$ to $\\phif(z=0.3) = 0.49$ and the characteristic mass of the low-mass component increasing from $\\Mstarf(z=0.5) = 9.41$ to $\\Mstarf(z=0.3) = 9.54$. This evolution in faint red galaxies is nearly perfectly mirrored by a decrease in number of faint blue galaxies. As we have noted above, the faint end slope of the active and the passive population is consistent with being equal. The similarity of the steep faint-end slopes of the active and faint passive populations and their reciprocal change in number density is suggestive of the latter originating from the former by shutting off star formation. It is very tempting to identify these two faint galaxy populations sharing the same steep faint-end slope with Sm/Im/dIrr galaxies, and faint spheroidal galaxies, respectively. It is established that passively evolving dwarf galaxies cluster around massive galaxies \\citep{Zehavi+2005,Haines+2006,Haines+2007,Carlberg+2009}, and tidal interactions or ram pressure stripping may well lead to quenching and some subsequent phase mixing turning Sm/Im galaxies into faint spheroidal galaxies \\citep{Mayer+2001,Grebel+2003,Haines+2007}. Also, \\citet{McCracken+2008} finds that in the CFHTLS in the redshift bin $0.2 < z < 0.6$ the clustering amplitude for faint red galaxies is actually higher than that of bright red galaxies, which is to be expected in our interpretation. Following on previous studies of the stellar mass function in the COSMOS field \\citep{Ilbert+2009, Pozzetti+2009, Bolzonella+2009}, we present a new analysis of this data set that provides mass functions to fainter limits than has been previously probed at $z \\lesssim 1$. The resulting increase in dynamic range allows us to characterize and study features in the shape of the stellar mass function that deviate from a single Schechter function. We have tested whether these features could be introduced by a variety of systematic effects including both catastrophic photometric redshift errors as well as increasing photo-$z$ uncertainty at low masses, differences in the way stellar population models account for TP-AGB stars, and the ability of stellar mass codes to convert from luminosity to mass. We conclude that our results are robust to these effects, although the data are still limited by cosmic variance. Our key results follow: \\begin{itemize} \\item Neither the total nor the red (passive) or blue (star-forming) galaxy stellar mass functions can be well fit with a single Schechter function once the the mass completeness limit of the sample probes below roughly $3 \\times 10^{9} \\Msun$. We model this more complicated behavior using a double Schechter function with 6 free parameters, and present the fitting results for 4 redshift bins to $z=1$. \\item The bimodal nature of the mass function is {\\em not} solely a result of the blue/red dichotomy. Indeed, the blue mass function is already bimodal at $z \\sim 1$. This suggests a new 2-component model for galaxy formation that predates the appearance of the red sequence. \\item We propose two interpretations for this bimodal distribution, focusing on the ``dip'' in the blue mass function at $\\sim$10$^{10} \\Msun$. If the gas fraction increases at lower stellar masses, galaxies with $M_{\\rm baryon} \\sim 10^{10} \\Msun$ would shift to lower stellar masses, creating the observed dip. This would indicate a change in SFR efficiency, perhaps arising from the influence of supernovae feedback which likely sets in below scales of $\\sim$10$^{10} \\Msun$. In this picture, the baryonic mass function should not show a dip. Using published (cold) gas fractions as a function of stellar mass, we show that cold gas alone is not sufficient to eliminate the dip in the baryonic mass function, but that the addition of the hard to detect warm gas could potentially flatten out the baryonic mass function considerably. \\item Alternatively, the dip could be created by an enhancement of the galaxy assembly rate at $\\sim$10$^{11} \\Msun$, a phenomenon that naturally arises if the baryon fraction peaks at $M_{\\rm halo} \\sim 10^{12} \\Msun$. In this scenario we would identify the galaxies occupying the bump around $\\Mstar$ with central galaxies and the increasing fraction of satellite galaxies at lower mass with the second, fainter, component of the mass function and in particular the steep faint-end slope. \\item The low-mass end of the blue and total mass functions exhibit a steeper slope than has been detected in previous work. This can be interpreted as a {\\em steepening} slope, one that may increasingly approach the halo mass function value of -2. \\item While the dip feature is apparent in the total mass function at all redshifts, it appears to shift from the blue to red population, likely as a result of transforming high-mass blue galaxies into red ones. \\item At the same time, we detect a drastic upturn in the number of low mass red galaxies. Their increase with time seems to reflect a decrease in the number blue systems and so we tentatively associate them with satellite (dwarf spheroidal) galaxies that have undergone quenching due to environmental processes. \\end{itemize} While the broad dynamic range of the COSMOS data set allows us to begin to characterize some of the more subtle features of the stellar mass function, the single COSMOS field is still limited by cosmic variance. Future work over several fields will verify the trends here and shed light on possible interpretations by constraining how these features evolve with redshift." }, "0910/0910.5666_arXiv.txt": { "abstract": "We present the first X-ray analysis of the diffuse hot ionized gas and the point sources in IC131, after NGC604 the second most X-ray luminous giant \\hii\\ region in \\m. The X-ray emission is detected only in the south eastern part of IC131 (named IC131-se) and is limited to an elliptical region of $\\sim$200 pc in extent. This region appears to be confined towards the west by a hemispherical shell of warm ionized gas and only fills about half that volume. Although the corresponding X-ray spectrum has 1215 counts, it cannot conclusively be told whether the extended X-ray emission is thermal, non-thermal, or a combination of both. A thermal plasma model of $kT_e$\\,=\\,4.3\\,keV or a single power law of $\\Gamma$\\,$\\simeq$\\,2.1 fit the spectrum equally well. If the spectrum is purely thermal (non-thermal), the total unabsorbed X-ray luminosity in the 0.35\\,--\\,8\\,keV energy band amounts to $L_X=6.8\\ (8.7)\\times 10^{35}$\\,\\ergs. Among other known \\hii\\ regions IC131-se seems to be extreme regarding the combination of its large extent of the X-ray plasma, the lack of massive O stars, its unusually high electron temperature (if thermal), and the large fraction of $L_X$ emitted above 2\\,keV ($\\sim$40\\,--53\\%). A thermal plasma of $\\sim$4\\,keV poses serious challenges to theoretical models, as it is not clear how high electron temperatures can be produced in \\hii\\ regions in view of mass-proportional and collisionless heating. If the gas is non-thermal or has non-thermal contributions, synchrotron emission would clearly dominate over inverse Compton emission. It is not clear if the same mechanisms which create non-thermal X-rays or accelerate CRs in SNRs can be applied to much larger scales of 200\\,pc. In both cases the existing theoretical models for giant \\hii\\ regions and superbubbles do not explain the hardness and extent of the X-ray emission in IC131-se. We also detect a variable source candidate in IC131. It seems that this object is a high mass X-ray binary whose optical counterpart is a B2-type star with a mass of $\\sim$9$M_{\\odot}$. ", "introduction": "This study is part of the \\cxo\\ ACIS Survey of M33 \\citep[\\chase,][]{plu08} and presents the first deep, high resolution X-ray images of \\object{IC131}. Within \\chase's sensitivity limit of $\\sim$$2\\times 10^{34}$\\, erg\\,s$^{-1}$ (0.35\\,--\\,8\\,keV) IC131 is, along with \\object{NGC604} \\citep{tull08}, the only giant \\hii\\ region (GHR) in \\object{\\m}\\ which shows significant diffuse extended X-ray emission. Because of their large sizes, masses, stellar contents, and luminosities, GHRs \\citep[cf.][]{ken84} provide excellent laboratories for studying the relation between stellar feedback and the evolution of the individual components of the ISM inside them. GHRs like 30\\,Dor \\citep{wang99,town06}, N51D \\citep{coop04}, or NGC604 \\citep{tull08} are strong X-ray emitters thanks to their massive O star population. The heavy mass loss of these stars can produce strong colliding winds which, together with contributions from SNe and SNRs, shock-heat the gas and forces it to emit thermal X-rays at temperatures around $kT_e=0.6\\pm0.2$\\,keV. In some cases the X-ray spectra of \\hii\\ regions in the Milky Way and the LMC require, in addition to a thermal component, also a non-thermal component \\citep[e.g.,][]{wolk02,bam04,coop04,smith04,muno06,mad09}. The non-thermal radiation could be explained by synchrotron emission, inverse Compton scattering, or by particle acceleration in shock regions \\citep{par04,arsch08}. In all cases, the X-ray emission could be linked to massive stars or their successors. The issue one faces with GHRs is to find a mechanism which can ionize the gas on spatial scales which can easily exceed that of normal \\hii\\ regions by one to two orders of magnitude and is consistent with the stellar energy feedback, X-ray luminosity, gas mass, and other predictable quantities. Although IC131 has been observed in numerous M33 surveys across almost all wavelengths \\citep[e.g.,][]{rosa84,land92,long96,hippe03,pietsch04,mass06,tab07,plu08}, little is known about the stellar population, its age, and the different components of the ISM. The \\chase\\ data set, supplemented by other observations at optical and infrared wavelengths, is the first data set that has the combination of S/N and spatial resolution to permit a detailed study of the X-ray emitting gas in IC131. Therefore, the primary focus of this work is to study the hot ionized medium (HIM) by constraining basic plasma parameters, to determine the likely ionization mechanism of the hot gas, and to compare properties of the gas to other star forming regions to learn more about the evolution of the gas in GHRs. The second purpose of this study focuses on the analysis of the point source population in IC131. In the following, we give an overview of the observations used in this study and the data reduction steps that have been applied (Section \\ref{sec2}). We then present high resolution (2$\\arcsec$) images of the diffuse X-ray emission in different energy bands and compare these to multi-wavelength data to constrain the morphology of IC131 (Section \\ref{spec_img}). A spectral analysis is carried out in Section \\ref{spec_res} while Section \\ref{spec_dis} discusses the possible origin and ionization mechanism of the HIM. Section \\ref{point_dis} deals with the analysis of the X-ray emitting point sources detected in IC131 and Section \\ref{con} gives a concluding summary of the most important results. ", "conclusions": "\\label{con} Compared to known superbubbles and \\hii\\ regions IC131-se appears to be peculiar with respect to its lopsided X-ray morphology, the large linear extent of the X-ray emission, the lack of massive O stars, its high electron temperature (in case the X-ray gas is thermal), and large fraction of hard X-rays. The X-ray spectrum of the extended emission in IC131-se can be equally well fitted by an absorbed power law ($\\Gamma \\simeq 2.1$) or an absorbed thermal plasma model ($kT_e\\simeq 4.3$\\,keV). These models predict a total (0.35\\,--\\,8.0\\,keV) unabsorbed X-ray luminosity of about $9\\times10^{35}$\\ergs\\ and $7\\times10^{35}$\\ergs, respectively, with 39\\% and 53\\% of the luminosity being emitted above 2\\,keV. Apparently IC131 possesses not only the hardest X-ray spectrum among the known GHRs, but, in case the gas is thermal, also the hottest X-ray plasma. Besides the facts that thermal X-ray emission is expected from GHRs and that there might be weak emission lines below 1\\,keV, the disagreement between the number of SNe progenitors obtained from the IMF (9\\,--\\,21) and that derived from the internal thermal energy (118$\\pm$59) argues against a pure thermal origin of the gas. We estimate a low electron density of $n_e\\simeq 0.05$\\,cm$^{-3}$ (assuming $f_X =0.9$) and an X-ray gas mass of about 3300\\,$M_{\\odot}$. The age of the stellar population inside the bubble seems to range from 8 to 77\\,Myr and given a cooling timescale of about 1\\,Gyr, the gas did not have time to cool significantly. If the gas is thermal, the standard bubble models require shocks from SNRs and SNe to heat the gas. The fundamental problem which argues against a thermal plasma is that we are not aware of a mechanism for GHRs which can produce electron temperatures anywhere close to $kT_e=4$\\,keV. Non-thermal X-ray emission is also known to originate from SNRs, GHRs or superbubbles and the spectral power law fit is statistically as good as the one for a thermal plasma. Synchrotron emission as opposed to inverse Compton scattering is a plausible mechanism for the assumed magnetic field strength of $B=20\\mu G$ and the total energy of the relativistic synchrotron-emitting electron population is about 14\\% to 30\\% of the total energy provided by SNe. If a non-thermal component exists, synchrotron losses clearly dominate over losses from inverse Compton scattering. For a purely non-thermal origin, however, it remains to be investigated what kinds of mechanisms are able to produce non-thermal X-rays or accelerate particles to relativistic energies on scales of 200\\,pc. To allow for a more quantitative analysis of the non-thermal X-ray emission, models are required which predict the fraction of non-thermal radiation from GHRs. A combined thermal and non-thermal model seems to be generally appropriate, too, irrespective of the unreasonably high column density and X-ray luminosity derived from those fits. Unfortunately, with the present data the nature of the extended X-ray emission cannot be conclusively determined. Clearly, IC131-se is an important object and challenges the standard bubble model as well as our understanding of CR acceleration in superbubbles. It is remarkable and remains to be understood why objects similar to IC131-se have not been detected before. Future investigations would greatly profit from more sensitive and deeper X-ray observations (as for example provided by IXO), high-resolution optical spectrophotometric data, and models which can explain electron temperatures $>$\\,1\\,keV and make predictions on the non-thermal X-ray emission in GHRs. The deeper X-ray observations could establish the presence of X-ray emission lines and settle the question on the nature of the emission, while a detailed stellar population analysis could provide a more accurate IMF. Clearly, NGC604 and IC131-se seem to be in completely different evolutionary stages. All O-type stars in IC131-se seem to have exploded as SNe, whereas the western part of NGC604 awaits the first SNe to occur. We detect only one X-ray point source in IC131. This source, CXO\\,J013315.10+304453.0 (FL073), appears to be time variable and is possibly a HMXRB with an optical counterpart which could be a B2V star with a mass of about 9M$_{\\odot}$." }, "0910/0910.2270_arXiv.txt": { "abstract": "Spectral line and continuum observations of the ionized and molecular gas in G20.08-0.14 N explore the dynamics of accretion over a range of spatial scales in this massive star-forming region. Very Large Array observations of NH$_3$ at $4\"$ angular resolution show a large-scale (0.5 pc) molecular accretion flow around and into a star cluster with three small, bright \\HII regions. Higher resolution ($0.4''$) observations with the Submillimeter Array in hot core molecules (CH$_3$CN, OCS, and SO$_2$) and the VLA in NH$_3$, show that the two brightest and smallest \\HII regions are themselves surrounded by smaller scale (0.05 pc) accretion flows. The axes of rotation of the large and small scale flows are aligned, and the timescale for the contraction of the cloud is short enough, 0.1 Myr, for the large-scale accretion flow to deliver significant mass to the smaller scales within the star formation timescale. The flow structure appears to be continuous and hierarchical from larger to smaller scales. Millimeter radio recombination line (RRL) observations at $0.4\\arcsec$ angular resolution indicate rotation and outflow of the ionized gas within the brightest \\HII region (A). The broad recombination lines and a continuum spectral energy distribution (SED) that rises continuously from cm to mm wavelengths, are both characteristic of the class of \\HII regions known as ``broad recombination line objects''. The SED indicates a density gradient inside this \\HII region, and the RRLs suggest supersonic flows. These observations are consistent with photoevaporation of the inner part of the rotationally flattened molecular accretion flow. We also report the serendipitous detection of a new NH$_3$ (3,3) maser. ", "introduction": "\\label{intro} Massive star-forming regions (MSFRs) with O stars are usually identified by a group of hypercompact (HC) \\HII or ultracompact (UC) \\HII regions found together, deeply embedded in a dense molecular cloud \\citep[][]{GL09,Church02,Hoar07}. That several \\HII regions are typically found within each star-forming region indicates that massive stars form together in small clusters. Furthermore, the infrared luminosity and radio continuum brightness of the individual \\HII regions suggest that some of them may themselves contain more than one massive star. Thus, the spatial structure of massive star-forming regions is clustered and hierarchical: the star-forming regions contain a number of separate HC and UC \\HII regions, each of which may in turn contain a few massive stars. Low angular resolution, single-dish, molecular line surveys of MSFRs show evidence for large scale contraction of the embedding molecular clouds \\citep{WE03,KW08}. Higher angular resolution observations of some of these regions identify velocity gradients consistent with rotation and inflow. In addition to the accretion flows seen on the large-scale ($\\sim 0.3-1$ pc) of the embedding molecular cloud (G10.6-0.4: Ho \\& Haschick 1986, Keto et al. 1987a, Keto 1990; G29.96-0.02: Olmi et al. 2003), accretion flows are also seen on smaller ($\\leq 0.1$ pc) scales around individual HC and UC \\HII regions (G10.6-0.4: Keto et al. 1988, Sollins et al. 2005a; W3(OH): Keto et al. 1987b, Keto et al. 1995; W51e2: Zhang \\& Ho 1997, Young et al. 1998; G28.20-0.05: Sollins et al. 2005b; G24.78+0.08: Beltr\\'an et al. 2004, 2005, 2006, Galv\\'an-Madrid et al. 2008; G29.96-0.02: Beuther et al. 2007). It is unclear how the flows on different length scales are related. In the case of G10.6-0.4, the cluster-scale accretion flow can be traced down from the largest cloud scale to the small scale of the brightest \\HII region, but it is not known whether this holds for other objects. For example, in a survey of MSFRs, selected on the basis of IRAS colors and specifically excluding those with \\HII regions, multiple bipolar molecular outflows (implying the presence of accretion flows) are seen in random orientations \\citep{Beu02a,Beu02,Beu03}. The different orientations of these smaller-scale flows suggest separate, individual centers of collapse. This comparison raises the question whether a large-scale coherent flow is required for the formation of the most massive stars, O stars ($M_\\star > 20$ $\\Msun$) capable of producing bright \\HII regions, whereas B stars require only smaller scale flows. It is also unclear what happens in an accretion flow when the inflowing molecular gas reaches the boundary of an embedded \\HII region. Previous observations suggest that the \\HII regions in an MSFR that are surrounded by accretion flows, may be best understood as deriving from the continuous ionization of the accretion flow \\citep{Keto02, Keto03, Keto07}, rather than as a dynamically separate expanding bubble of ionized gas within the flow. Part of the ionized gas may continue to the central star or stars and part escapes off the rotationally flattened accretion flow as a photoevaporative outflow perpendicular to the plane of rotation \\citep{Hol94, YW96, Liz96, John98, Lugo04}. The outflow is accelerated to supersonic speeds by the density gradient maintained by the stellar gravity \\citep{Keto07}. Because the extent of an ionized outflow is generally larger than the region of ionized inflow, in most cases the outflow should be detected more easily than the inflow. \\HII regions classified as ``broad recombination line objects\" (BRLO) \\citep{JaffMP99,Sewi04} show steep density gradients and supersonic flows \\citep{KZK08}, consistent with photoevaporation and acceleration. It is not known whether all BRLO are associated with accretion. If the accretion surrounding an O star cluster is continuous from the largest to the smallest scales, this must be the case. There are only a handful of radio recombination line (RRL) observations that spatially resolve the ionized flow within an HC \\HII region. Velocity gradients consistent with outflow and rotation in the ionized gas have been previously reported for W3(OH) \\citep{Keto95}, W51e2 \\citep{KK08}, and G28.20-0.05 \\citep{Sewi08}. Observations of the very massive and spatially large G10.6-0.4 \\HII region made at the VLA in the highest possible angular resolution are able to map the inflowing ionized gas \\citep{KW06}. In order to study the accretion dynamics over a range of scales in a MSFR, from the cluster scale down to the scale of individual HC \\HII regions and within the ionized gas, we set up a program of radio frequency molecular line, recombination line, and continuum observations at two telescopes and with several different angular resolutions. For this study we chose the massive star formation region G20.08-0.14 North (hereafter G20.08N), identified by three UC and HC \\HII regions detected in the cm continuum by \\cite{WC89}. The total luminosity of the region is $L \\sim 6.6 \\times 10^5$ $L_\\odot$ for a distance of 12.3 kpc.\\footnote{ Both near and far kinematic distances have been reported for G20.08N. The near value given by \\cite{Dow80} ($d \\approx 4.1$ kpc) is the most commonly quoted in the previous literature. In contrast, \\cite{Fish03} and \\cite{AnBan09} report that this region is at the far kinematic distance ($d \\approx 12.3$ kpc). We will assume the far distance throughout the rest of the paper. For reference, a scale of $0.5\\arcsec$ corresponds to $\\approx 6000$ AU (0.03 pc). The total luminosity of the region was estimated to be $L \\sim 7.3 \\times 10^4$ $L_\\odot$ assuming the near kinematic distance \\citep{WC89}. Correcting for the location at the far distance, the luminosity is $L \\sim 6.6 \\times 10^5$ $L_\\odot$.} Previous observations suggest accretion in the G20.08N cluster. Molecular-line observations show dense gas embedding the \\HII regions \\citep{Tur79,Plume92}. Molecular masers, generally associated with ongoing massive-star formation, have been detected in a number of studies (OH: Ho et al. 1983; H$_2$O: Hofner \\& Churchwell 1996; and CH$_3$OH: Walsh et al. 1998). \\citet{KW07,KW08} observed large-scale inward motions consistent with an overall contraction of the embedding molecular cloud. Those authors also observed SiO line profiles suggestive of massive molecular outflows, further evidence for accretion and star formation. The recombination line spectra show broad lines \\citep{Gar85, Sewi04}, presumably due to large, organized motions in the ionized gas. However, the previous observations do not have the angular resolution and the range of spatial scales needed to confirm the presence of accretion flows and study them in detail. In this paper we report on several observations of G20.08N and discuss our findings. We confirm active accretion within the cluster. Furthermore, we find that the parsec-scale accretion flow fragments into smaller flows around the individual HC \\HII regions, and that the gas probably flows from the largest scale down to the smallest scale. This continuous and hierarchical accretion may be necessary to supply enough mass to the small-scale flows to form O-type stars, in contrast to low- and intermediate-mass star-forming regions with stars no more massive than $\\sim 20$ $\\Msun$, where isolated accretion flows around individual protostars may be sufficient. ", "conclusions": "\\label{conclu} We report radio and mm observations of the molecular and ionized gas toward the O-star cluster G20.08N, made with an angular resolution from $\\sim 0.1$ pc to $\\sim 0.01$ pc. Our main findings can be summarized as follows: \\begin{enumerate} \\item We find a large-scale ($\\sim 0.5$ pc) accretion flow around and into a star cluster with several O-type stars, identified by one UC and two HC \\HII regions. This flow is rotating and infalling towards its center. The two HC \\HII regions are surrounded by smaller accretion flows ($\\sim 0.05$ pc), each of them with the signature of infall too. The brightest (toward \\HII region A) is detected in mm emission lines, and rotates in concordance with the large-scale flow. \\item The similar orientations of the flows at small and large scales, as well as their dynamical timescales ($\\sim 10^4$ yrs and $\\sim 10^5$ yrs respectively), and masses ($\\sim 10~\\Msun$ and $\\sim 10^3~\\Msun$ respectively), suggest that, if O stars are forming in G20.08N (as it is observed), then the smaller scales ought to be resupplied from the larger scales. The same result has been found in recent numerical simulations of massive star formation in clusters. \\item The brightest HC \\HII region (A) has a rising SED from cm to mm wavelengths and broad hydrogen recombination lines. Both characteristics suggest density gradients and supersonic flows inside the \\HII region. A tentative velocity gradient is detected in the recombination line emission of this source, suggesting rotation and outflow in the ionized gas at the innermost scales. \\HII region A can be interpreted as the inner part of the surrounding molecular accretion flow, with the observed ionization being produced by photoevaporation. \\end{enumerate}" }, "0910/0910.0275_arXiv.txt": { "abstract": "We present blue optical spectra of 92 members of $h$ and $\\chi$ Per obtained with the WIYN telescope at Kitt Peak National Observatory. From these spectra, several stellar parameters were measured for the B type stars, including $V$ sin $i$, $T_{\\rm eff}$, log $g_{\\rm polar}$, $M_{\\star}$, and $R_{\\star}$. Str\\\"{o}mgren photometry was used to measure $T_{\\rm eff}$ and log $g_{\\rm polar}$ for the Be stars. We also analyze photometric data of cluster members and discuss the near-to-mid IR excesses of Be stars. ", "introduction": "NGC 869 and NGC 884 ($h$ and $\\chi$ Persei, respectively) are a well known double cluster rich in massive B-type stars, and have been the focus of many studies over the years. Recent studies show that NGC 869 and NGC 884 have nearly identical ages of $\\sim$ 13--14 Myr, common distance moduli of dM $\\sim$ 11.85, and common reddenings of E(B-V) $\\sim$ 0.5--0.55 (\\cite[Currie et al.\\ 2009]{Currie_etal09}, \\cite[Slesnick et al.\\ 2002]{Slesnick_etal02}, \\cite[Bragg \\& Kenyon 2005]{Bragg_etal05}). \\cite[Currie et al.\\ (2008; hereafter C08)]{Currie_etal08} identified two populations of NGC 869 and NGC 884 stars with detected Spitzer MIPS-24 $\\mu$m excess emission: 20 A and F-type stars with luminous debris disk emission and 57 brighter, earlier stars with weaker excess emission. They identify most of the latter group as candidate Be stars. However, only 21 were previously listed as Be stars (eg. \\cite[Bragg \\& Kenyon 2002]{Bragg_etal02}, \\cite[Slesnick et al.\\ 2002]{Slesnick_etal02}). In this study, we analyze blue optical spectra of 92 early-type cluster members, including 16 candidate Be stars from C08, and investigate their near-to-mid infrared (IR) excesses. With continued monitoring of these stars in the both the optical and IR regimes, we hope to explore these excesses as a reasonable means for identifying potential Be stars within clusters, as well as to investigate the transient natures of the disks surrounding the known Be stars in NGC 869 and NGC 884. \\vspace{-0.1in} ", "conclusions": "We have measured the physical parameters of 77 B-type stars and 15 Be stars in NGC 896 and NGC 884. Sixteen Be candidates from C08 are present in our sample or that of HG06. Of these 16 Be candidates, 3 stars show no evidence of emission in our optical data and are likely transient Be stars. Ten of these Be candidates do show emission in our spectra. Those Be candidates without emission in our spectra should be monitored in the future to further investigate their transient nature. In the future, IRAC 3.6-5.8 $\\mu$m data will be combined with the optical and IR fluxes used here to investigate the observed SEDs. We will fit the new SEDs using modern flux models rather than blackbody curves. Modifications accounting for variable reddening throughout the clusters will also be made. These new SED fits can then be used to model the Be disk sizes and temperatures." }, "0910/0910.2877_arXiv.txt": { "abstract": "Current X-ray observatories make it possible to follow the evolution of transient and variable X-ray binaries across a broad range in luminosity and source behavior. In such studies, it can be unclear whether evolution in the low energy portion of the spectrum should be attributed to evolution in the source, or instead to evolution in neutral photoelectric absorption. Dispersive spectrometers make it possible to address this problem. We have analyzed a small but diverse set of X-ray binaries observed with the {\\it Chandra} High Energy Transmission Grating Spectrometer across a range in luminosity and different spectral states. The column density in individual photoelectric absorption edges remains constant with luminosity, both within and across source spectral states. This finding suggests that absorption in the interstellar medium strongly dominates the neutral column density observed in spectra of X-ray binaries. Consequently, evolution in the low energy spectrum of X-ray binaries should properly be attributed to evolution in the source spectrum. We discuss our results in the context of X-ray binary spectroscopy with current and future X-ray missions. ", "introduction": "X-ray binaries show a remarkable variety of extreme phenomena, including rapid variability in many wavelength bands, the production of relativistic jets and hot winds, and outbursts that span factors of $10^{6}$ in luminosity (for a review of black hole transients, see Remillard \\& McClintock 2006). Monitoring observations with the {\\it Rossi X-ray Timing Explorer} and {\\it Swift} enable the evolution of X-ray spectral and timing properties of X-ray binaries to be traced in order to understand the accretion process (see, e.g., Rykoff et al.\\ 2007). Model degeneracies that may occur in fitting a single spectrum can often be resolved by examining how multiple spectra evolve with time. Major changes within the outbursts of transient and persistent X-ray binaries are typically linked with changes in the nature of the accretion disk (see, e.g., Esin, McClintock, \\& Narayan 1997; Homan et al.\\ 2001; also see Miller et al.\\ 2006a). Tracking accretion disk evolution is therefore a primary goal of monitoring observations. The disk is directly visible in soft X-rays in black hole X-ray binaries, and likely also in neutron star binaries (e.g. Cackett et al.\\ 2008a). However, the K shell photoelectric absorption edges from most abundant elements fall below 2~keV, as do Fe L shell edges. This absorption can complicate efforts to study disk properties, especially when using a spectrometer with limited energy range and resolution. Prominent theoretical treatments of accretion flows predict modest winds (e.g. Begelman, McKee, \\& Shields 1983) and/or outflows with a high ionization parameter (e.g. Narayan \\& Raymond 1999). Observationally, even the most extreme winds detected in X-ray binaries do not appear to generate enough column density to add appreciably to the neutral absorption column (Miller et al.\\ 2006). Nevertheless, it is not uncommon to allow the equivalent neutral hydrogen column density to vary in modeling multiple spectra of an X-ray binary (see, e.g. Brocksopp et al.\\ 2006; Caballero-Garcia et al.\\ 2009; Cabanac et al.\\ 2009). Even in cases where the neutral column density is held constant or jointly determined, this method is adopted somewhat arbitrarily, without specific reference to an observational or theoretical basis (see, e.g., Sobczak et al.\\ 2000; Miller et al.\\ 2001; Nowak, Wilms, \\& Dove 2002; Park et al.\\ 2004; Gierlinski \\& Done 2004; Kalemci et al.\\ 2005; Rykoff et al.\\ 2007). Dispersive X-ray spectroscopy can separate variations in absorption from variations in the source spectrum. Even the resolution afforded by CCD spectrometers blends many weak edges into the continuum, and the optical depth and wavelength of stronger edges can be difficult to measure precisely. For instance, in fits to CCD spectra, the multiple Fe L edges can be modeled satisfactorily in terms of a single step function. In constrast, {\\it Chandra} High Energy Transmission Grating Spectrometer observations of X-ray binaries have resolved the individual Fe L edges and their detailed properties (see, e.g., Schulz et al. 2002, Juett et al.\\ 2006). Beyond simple edges, dispersive X-ray spectroscopy has clearly revealed the neutral O I 1s-2p resonance absorption line that was first glimpsed with the crystal spectrometer aboard {\\it Einstein} (Schattenberg \\& Canizares 1986, Paerels et al.\\ 2001). In this Letter, we extend recent work using high resolution X-ray spectroscopy by examining the evolution of neutral absorption in a sample of X-ray binaries. We find no evidence for variable absorption, consistent with an interstellar origin for neutral absorption. ", "conclusions": "To better understand the evolution of the low energy spectrum of X-ray binaries, we made fits to individual photoelectric absorption edges in high resolution X-ray spectra of selected sources. The column density measured in individual edges is not observed to vary across different spectral states, nor over a broad range in luminosity (see Table 1 and Figure 2). This suggests that gas from X-ray binaries is not typically an important source of the neutral absorption observed in the spectra of these systems. Rather, neutral absorption must be dominated by the ISM. A similar conclusion was reached by Juett, Schulz, \\& Chakrabarty (2004) based on upper limits on the velocity dispersion of the ISM as measured through absorption lines. Evolution in the low energy spectrum of typical X-ray binaries, then, is best attributed to evolution in the source continuum. Neutral absorption in X-ray spectra is often fit by a single model that parameterizes the accumulated absorption from individual edges as an equivalent neutral hydrogen column density. Values obtained from high resolution spectra are likely to give the best measure of a true equivalent total column density. In practice, instrumental problems such as internal scattering, carbon build-up from optical blocking filters, and gain drift could prevent the adoption of a gratings-derived value for the neutral column. In such cases, our results suggest that a single value of the equivalent neutral hydrogen column density should be used to fit multiple spectra from monitoring observations of a given source with a given detector. The outflows that are observed in X-ray binaries are highly ionized -- dominated by He-like and H-like charge states (see, e.g., Lee et al.\\ 2002, Miller et al.\\ 2004, Miller et al.\\ 2006b, Schulz et al.\\ 2008). An especially dense wind was observed in GRO J1655$-$40, and even in that case the ionized columns observed are insufficient to create strong absorption edges that could be mistaken for additional neutral absorption (Miller et al.\\ 2006c, 2008). Moreover, in sources such as H 1743$-$322, GRO J1655$-$40, and GRS 1915$+$105, a paradigm is emerging wherein ionized winds are active in soft, disk-dominated states, but absent in spectrally hard states like those that typically hold when sources accrete at a low fraction of their Eddington limit (Miller et al.\\ 2006b, 2006c, Miller et al.\\ 2008, Neilsen \\& Lee 2009). In Figure 2, it is clear that any variation in ionized winds across states does not affect the column density measured in neutral edges. Our results are based on spectra which only reach down to approximately $0.01~{\\rm L}_{\\rm Edd.}$. Sensitive high-resolution X-ray spectra that would permit strong constraints on absorption variability have not yet been obtained from sources at lower luminosity. However, theoretical arguments again point to ionized outflows that would contribute little to a neutral column. Winds from advection-dominated accretion flows are expected to be extremely hot (since advective flows are very hot), and line spectra should be dominated by He-like and H-like charge states (e.g. Narayan \\& Raymond 1999). Recent observations of the stellar-mass black hole V404 Cyg, which accretes at about $10^{-5}~ {\\rm L}_{\\rm Edd.}$, are able to rule-out the winds predicted by some advective models (Bradley et al.\\ 2007). This further suggests that ionized outflows are not likely to contribute significantly to line-of-sight absorption, even as sources approach quiescence. There are at least two classes of sources where our results may not hold in all circumstances. Neutron star X-ray binaries known as ``dippers'' are viewed at high inclinations, and material in the outer disk can block emission from the central engine (see, e.g., Diaz Trigo et al.\\ 2006). Within such flux dips, the observed neutral absorption column may vary due to the changing geometry within the binary. Similarly, winds from massive stars are known to be clumpy and to sometimes cause flux dips; some of these dips may also cause variations in the observed equivalent neutral hydrogen column density (e.g. Balucinska-Church et al.\\ 2000). \\vspace{0.1in} \\noindent We thank Joern Wilms for creating the ``tbnew'' model used in this work, and for guidance on its implementation. We acknowledge helpful comments from the anonymous referee that improved this paper. J.M.M. gratefully acknowledges funding from the {\\it Chandra} Guest Observer program. EMC gratefully acknowledges support provided by NASA through the Chandra Fellowship Program, grant number PF8-90052. RCR acknowledges STFC for financial support." }, "0910/0910.0043_arXiv.txt": { "abstract": "{} {We are studying the dust properties of four low metallicity galaxies by modelling their spectral energy distributions. This modelling enables us to constrain the dust properties such as the mass, the temperature or the composition to characterise the global ISM properties in dwarf galaxies. } {We present 870 \\mic\\ images of four low metallicity galaxies (NGC~1705, Haro~11, Mrk~1089 and UM~311) observed with the Large \\APEX\\ BOlometer CAmera (LABOCA) on the Atacama Pathfinder EXperiment (\\APEX) telescope. We model their spectral energy distributions combining the submm observations of \\lab, \\twomass, \\iras, \\spitz\\ photometric data and the IRS data for Haro~11.} {We find that the PAH mass abundance is very low in these galaxies, 5 to 50 times lower than the PAH mass fraction of our Galaxy. We also find that a significant mass of dust is revealed when using submm constraints compared to that measured with only mid-IR to far-IR observations extending only to 160 \\mic. For NGC~1705 and Haro~11, an excess in submillimeter wavelengths is detected when we use our standard dust SED model. We rerun our SED procedure adding a cold dust component (10 K) to better describe the high 870 \\mic\\ flux derived from LABOCA observations, which significantly improves the fit. We find that at least 70 $\\%$ of the dust mass of these two galaxies can reside in a cold dust component. We also show that the subsequent dust-to-gas mass ratios, considering HI and CO observations, can be strikingly high for Haro~11 in comparison with what is usually expected for these low-metallicity environments. Furthermore, we derive the star formation rate of our galaxies and compare them to the Schmidt law. Haro~11 falls anomalously far from the Schmidt relation. These results may suggest that a reservoir of hidden gas could be present in molecular form not traced by the current CO observations. While there can be a significant cold dust mass found in Haro~11, the SED peaks at exceptionally short wavelengths (36 \\mic), also highlighting the importance of the much warmer dust component heated by the massive star clusters in Haro~11. We also derive the total IR luminosities derived from our models and compare them with relations that derive this luminosity from \\spitz\\ bands. We find that the Draine $\\&$ Li (2007) formula compares well to our direct IR determinations. } {} ", "introduction": "The understanding of the evolution of a galaxy requires knowledge of the roles of the different actors controlling the evolution of the Interstellar Medium (ISM) and the subsequent feedback on star formation activity. Despite its low fraction of the total mass of a galaxy (less than 1$\\%$), dust plays a prominent role in the heating and cooling of the ISM and thus tightly influences the overall physics of a galaxy. Since dust absorbs the stellar radiation and reemits it in a wide range of wavelengths, the star formation rate (SFR) as well as other fundamental parameters of a galaxy, such as its age, can be indirectly studied through the dust emission itself. {\\revisedbis The Spectral Energy Distribution (SED) of a galaxy is its spectral footprint from which we can study the physical processes taking place in the galaxy since it synthesises the contribution of all its components to the emission of the galaxy. Using this tool, we can peer into the window of the integrated history of the galaxy and disentangle the various physical actors (stars, HII regions, molecular clouds) and processes (stellar radiation, dust emission) involved~\\citep[][see also $\\S$ 5 of this paper]{Draine2007,Galliano_Dwek_Chanial_2008}.} While dust hinders the interpretation of ultraviolet (UV) and optical wavelengths, in the Mid Infrared (MIR), Far Infrared (FIR) and submillimetre (submm) wavelengths, dust emission and absorption properties expose different physical environments, from the most vigorous star formation and AGN activity \\citep[e.g.][]{Gordon1995, Wu2007} to the more quiescent diffuse media \\citep{Bernard1996,Arendt1998}. Many processes linked to star formation such as stellar winds \\citep{Hoefner2009}, supernovae shocks, photodestruction by high-mass stars etc. can also affect the spatial distribution and the local properties and abundance of the different dust components of a galaxy such as Polycyclic Aromatic Hydrocarbons \\citep[PAHs,][]{OHalloran2006}, amorphous carbon grains, silicates or composite grains, manifesting themselves in the MIR to submm wavelengths. Studying the interplay between galaxy properties and metal enrichment is crucial to understand galaxy evolution. The metallicity of a galaxy is deeply linked with the dust properties of the ISM and its substructures such as HII regions and molecular clouds, but just how it affects the ISM is currently poorly known. Dwarf galaxies in the Local Universe, are metal-poor galaxies, and are thus convenient laboratories to study the effects of metallicity (Z) on the gas and dust. They exhibit a wide variety of physical conditions, and their star formation properties and ISM represent the closest analogs to proto-galaxies of the early universe. Indeed, dwarf galaxies are small and may compared to high redshift galaxies which also present lower metallicities \\citep{Lara_Lopez_2009}. They are also considered to be the building blocks of much larger and more metal-rich galaxies \\citep[Review by][]{Tosi2003}. They also show analogies with Gamma Ray Bursts (GRB) hosts whose ISM usually exhibit moderate chemical enrichment with a median metallicity of 1/10 \\zsun \\citep{Chen2009}. They finally show evidence for older stellar populations than their metallicity suggests \\citep[e.g.][]{Aloisi1998}, posing enigmatic issues for galaxy evolution models. Many studies have been carried out to grasp this apparent paradox. \\citet{Lisenfeld_Ferrara_1998} confirmed that the dependence of the dust-to-gas mass ratio (D/G) in low metallicity galaxies was a function of metallicity using \\iras\\ observations. \\citet{James2002}, \\citet{Walter2007} and \\citet{Hirashita2008} concluded likewise using \\jcmt/SCUBA submm, \\spitz\\ MIR/FIR and \\akari\\ (FIR) observations. Finally, \\citet{Galliano_Dwek_Chanial_2008} observed some systematic deviations between dust abundances of very low metallicity systems and what is expected for supernova-condensed dust. At MIR wavelengths, low metallicity systems also show prominent differences in the dust properties compared to the more metal-rich systems. For example, PAH features are strikingly diminished as metallicities drop \\citep[e.g.][]{Madden2005, Engelbracht2005, Wu2007, Engelbracht2008} compared to metal-rich galaxies, in spite of the role the smallest grains play in the energy balance of galaxies \\citep{Rubin2009}. Some studies suggest that PAH emission depends on the hardness or strengh of the illuminating radiation field \\citep{Madden2006, Engelbracht2008, Gordon2008, Bendo2008}. The consequence of lowering the metallicity of a galaxy is the decrease in dust opacity resulting in harder and stronger radiation fields. The dearth of PAHs in low metallicity galaxies has also been explained by the destructive effects of supernovae \\citep{OHalloran2006,OHalloran2008} or by the delayed injection of PAHs by AGB stars \\citep{Galliano_Dwek_Chanial_2008}. Broad wavelength coverage of the MIR to submm regime is imperative to constrain the modelling of the observed SEDs, leading to a better comprehension of the dust properties of galaxies. Since \\spitz\\ only observes dust emission at wavelengths shorter than 160 \\mic, submm data are necessary not only to enlarge the wavelength coverage at longer wavelengths to verify the dust models but also because the potential reservoir of cold grains ($\\le$ 15K), which contribute to this submm flux, may account for a significant amount of mass. Only a handful of galaxies of the Local Universe have been studied using submm ground-based instruments (e.g. \\jcmt/SCUBA). When submm observations of dwarf or late-type galaxies are studied, an excess in the dust SEDs is often found in the mm/submm domain \\citep{ Lisenfeld2001, Bottner2003, Dumke2004,Galliano2003, Galliano2005, Marleau2006, Bendo2006}. This excess can be interpreted as very cold dust ($\\le$10K), in which case more than 50$\\%$ of the total dust mass of these galaxies should reside in a very cold component. {\\revisedbis Cold dust is also needed to explain the break in the Gas to Dust mas ratio as a function of metallicity relaion for low-metallicity galaxies~\\citep{Galliano_Dwek_Chanial_2008, Munoz2009}}. The presence of this cold dust component is still a contentious issue in the ISM community and will have important consequences on our comprehension of ISM properties of low metallicity environments. \\citet{Lisenfeld2001}, \\citet{Reach1995}, \\citet{Dumke2004}, \\citet{Bendo2006} or \\citet{Meny2007} suggested that changes in dust emission properties (changes in dust emissivity or resonances related to dust impurities) should be responsible for boosting submm emission above the 15-20K thermal emission expected at these wavelengths. However, not all low metallicity galaxies show submm excess, as shown recently by the observations of the nearby Local Group Galaxy {\\revisedbis IC10 {\\revisedbis (Parkin et al. 2009 in prep)}, where the main} two star forming regions were isolated with ISO, \\spitz\\ and 850 \\mic\\ observations, the SEDs were modeled without invoking a very cold dust component. Moreover, \\citet{Draine2007} showed that their observations of mostly metal-rich galaxies can largely be reproduced by dust models which do not account for a very cold dust component, even in their limited number of cases where submm observations are present. Studies using submm observations for a wider range of metallicity values are necessary to check the relevance of these conclusions for low metallicity environments. In this paper, we present the first \\APEX/ \\lab\\ 870 \\mic\\ observations of dwarf galaxies: 1 extended galaxy (NGC~1705) and 3 compact sources (Haro~11, UM~311, Mrk~1089) (see Table~\\ref{Sample_properties}). We have combined these data with \\spitz\\ and/or \\iras\\ observations to produce global SEDs that we use to model the dust properties. The sample is small but covers a wide range of metallicities, from $\\sim$1/9 \\zsun\\ for Haro~11 to $\\sim$1/3 \\zsun\\ for NGC~1705. It also presents varied morphologies, size scales and characteristics: resolved or compact galaxies, disturbed and even merging environments. We describe the sample in $\\S$ 2 and the observations and data reduction in $\\S$ 3 and the images and photometry in $\\S$ 4. In $\\S$ 5, we present the SED modelling and discuss the results in $\\S$ 6. ", "conclusions": "The quantification of the dust mass of a galaxy aids our understanding of its evolution and star formation history. Larger dust masses are sometimes found in low metallicity galaxies when using submm constraints in the SED modelling. In this context, submm observations are clearly necessary to lead a more complete description of the distribution and properties of dust. We focused our paper on the dust modelling of four low metallicity galaxies observed with \\APEX/\\lab. \\\\ In this paper: \\begin{enumerate} \\item We present the first images of four dwarf galaxies carried out with the \\APEX/\\lab\\ instrument observing at 870 \\mic. \\item We construct the SEDs with \\spitz\\ IRAC and MIPS bands as well as the IRS spectra for Haro~11. We apply our SED model and determine the dust properties of these galaxies. \\item We find that the mass of PAHs accounts for 0.08 to 0.8 $\\%$ of the total dust mass of the galaxies, which is a factor of 5 to 50 lower than that of the Galaxy. \\item To investigate the influence of the submm constraints on the interpretation, we test the effect on the SED model results when submm 870 \\mic\\ observations are taken into account and compare with SED models not taking into account the 870 \\mic\\ flux, but with observational constraints at wavelengths only as long as 160 \\mic. We find that the use of submm observational constraints always leads to an increase of the total dust mass derived for our low metallicity galaxies. \\item We choose to include an additional component to account for the excess submm emission of NGC~1705 and Haro~11. A cold dust component ($\\sim$ 10K) with a $\\beta$ emissivity index of 1 substantially improves the fit. We find at least 70$\\%$ of the total dust mass residing in a cold ($\\sim$ 10K) dust component for these two galaxies. We note that describing a cold component of $\\beta$=2 does not give very different $\\bar{\\chi}$$^2$ values, but would give unrealistically larger D/G. Our results however do not rule out the hypothesis of a change in dust emissivity as a function of wavelength proposed in recent studies \\citep[e.g.][]{Dupac2003,Meny2007}. \\item While Haro~11 has a substantial ($\\sim$ 70$\\%$) cold dust component, it also harbours a significant fraction of dust mass (30$\\%$) in a warmer dust component ($>$ 25K). The SED peaks at unusually short wavelengts (36 \\mic), highlighting the importance of the warm dust. \\item We determine the D/G for Mrk~1089, NGC~1705 and Haro~11 to be 1.9 $\\times$ 10$^{-3}$, 4.1 $\\times$ 10$^{-3}$ and 0.2, respectively. For Mrk~1089 and NGC~1705, these D/G are consistent with current chemical evolution models. On the contrary, Haro~11 has an excessively high D/G considering the upper limits detected in HI and CO. Haro~11 also falls far from the Schmidt law, perhaps due to the observed deficit of gas in this galaxy. This could suggest the presence of a large amount of molecular gas. 10 times more molecular gas, compared to that deduced from CO measurements, may be present but not necessarily traced by CO observations. \\item From our SED models, we determine the total infrared luminosity of our galaxy sample to range from 5.8 $\\times$ 10$^{7}$ for NGC~1705 to 1.7 $\\times$ 10$^{11}$ for the LIRG Haro~11. These values of L$_{TIR}$ are systematically higher than those obtained using the \\citet{Dale_Helou_2002} formula but compare better to the \\citet{Draine_Li_2007} formula. While $\\sim$90$\\%$ of the dust mass is residing in the FIR to submm regime, not more than 6$\\%$ of the total IR luminosity in Haro~11 emerges from the FIR to submm (100 to 1100 \\mic), while most of the luminosity (70$\\%$) emerges in the NIR to MIR (3 \\mic\\ to 50 \\mic) window. This is in contrast to Mrk~1089 and NGC~1705 which distribute their luminosities more equally in these two wavelength windows. \\end{enumerate} Better observational coverage of the Rayleigh Jeans side of the SEDs should help us to disentangle possible scenarios to explain the excess detected in some galaxies in the submm: dominant diffuse ISM dust, modifications of the dust optical properties at submm wavelengths, very cold dust component etc. The \\hersc\\ guaranteed time key program of 50 dwarf galaxies (PI: S.Madden), which covers a wide range of metallicity values, will provide a broader coverage of wavelengths between 60 \\mic\\ to 600 \\mic\\ and will enable us to better sample the warm and cold dust. The observations will especially enable us to study the slope of the Rayleigh Jeans side of the SED and to learn if an independant very cold dust component does indeed exist. A following paper will discuss the systematic increase in the dust mass estimate with or without submm constraints and the implications for the D/G ratios for a broader sample of galaxies and study how it could possibly be influenced by metallicity. \\\\" }, "0910/0910.0004_arXiv.txt": { "abstract": "We use the overdensity field reconstructed in the volume of the COSMOS area to study the nonlinear biasing of the zCOSMOS galaxies. The galaxy overdensity field is reconstructed using the current sample of $\\sim$8500 accurate zCOSMOS redshifts at $I_{AB} < 22.5$ out to z$\\sim$1 on scales $R$ from 8 to 12 \\hh Mpc. By comparing the probability distribution function (PDF) of galaxy density contrast $\\delta_g$ to the lognormal approximation of the PDF of the mass density contrast $\\delta$, we obtain the mean biasing function $b(\\delta, z,R)$ between the galaxy and matter overdensity field and its second moments $\\hat{b}$ and $\\tilde{b}$ up to $z \\sim 1$. Over the redshift interval $0.41$, as values of $\\delta_g<-1$ are not possible. This implies that the galaxy bias $b$ must be a nonlinear function of $\\delta$ and, in general, it can also vary with redshift $z$, galaxy type and the smoothing scale $R$ used to define the density contrast: \\be b=b(\\delta, z, R). \\ee \\citet{Fry&Gaztanaga.1993} proposed to parameterise this function in terms of coefficients of the Taylor expansion \\be \\delta_g=b_0+b_1\\,\\delta+\\frac{b_2}{2}\\,\\delta^2+\\dots \\;, \\label{eq_biasfg} \\ee which are not fully independent as the conditions $\\langle \\delta_g \\rangle=0$ and $\\delta_g(\\delta=-1)=-1$ must hold. Galaxy biasing is also expected to have a stochastic element: for any given value of $\\delta$ there will be a whole distribution of values for $\\delta_g$. The stochasticity originates from a number of different sources. First, the dynamics of large scale flows depends on extra variables beyond the value of the local density contrast $\\delta$ (e.g. on the tidal tensor) and makes the bias relation nonlinear, non-local and stochastic \\citep{Catelan.etal.1998}. Second, the efficiency of galaxy formation depends on details of the gas physics. Third, galaxies are discrete objects and any attempt to reconstruct $\\delta_g$ will be effected by shot noise. \\citet{Dekel&Lahav.1999} have proposed a formalism which separately accounts for the nonlinearity and stochasticity of the biasing process. Galaxy biasing is described in terms of the conditional probability function $P(\\delta_g|\\delta)$ and its moments. A key quantity here is the mean biasing function $b(\\delta)$ defined by the conditional mean: \\be b(\\delta) \\delta = \\langle \\delta_g | \\delta \\rangle = \\int d \\delta_g P(\\delta_g|\\delta) \\delta_g . \\ee The mean biasing function $b(\\delta$) and its nonlinearity can be characterised by its second non-trivial moments: \\be \\hat{b} \\equiv \\frac {\\langle b(\\delta)\\delta^2 \\rangle}{\\sigma^2} \\ee and \\be \\tilde{b}^2 \\equiv \\frac {\\langle b^2(\\delta)\\delta^2 \\rangle}{\\sigma^2} , \\ee with $\\sigma^2$ the variance of the mass density contrast distribution. The parameter $\\hat{b}$ measures the slope of the linear regression of $\\delta_g$ against $\\delta$. In the case of linear biasing (see equation~\\ref{eq_biasdetlin}), both $\\hat{b}$ and $\\tilde{b}$ reduce to the constant bias. The ratio $\\tilde{b}/\\hat{b}$ is thus a measure of the nonlinearity in the biasing relation. Moreover, the local variance of $\\delta_g$ at fixed $\\delta$, $\\sigma^2_g(\\delta)$ can be used to quantify the degree of stochasticity of the biasing relation. Based on the Press-Schechter formalism and its extensions \\citep{Bond.etal.1991}, \\citet{Mo&White.1996} developed an analytical model for the mean biasing relation of the dark matter haloes. This assumes that large scale motions follow the spherical collapse approximation. The general case is discussed by \\citet{Catelan.etal.1998}. Related work has been presented in \\citet{Mo.etal.1997} and Porciani et al. (1998, see also Scannapieco \\& Barkana 2002) \\nocite{Porciani.etal.1998, Scannapieco&Barkana.2002} where two-point and higher-order statistics are considered. Following the analytical approach by \\citet{Mo&White.1996}, a number of studies based on N-body simulations were carried out to study the halo bias \\citep[e.g.][]{Jing.1998, Porciani.etal.1999, Sheth&Lemson.1999, Sheth&Tormen.1999, Jing.1999, Kravtsov&Klypin.1999, Sheth.etal.2001, Seljak&Warren.2004, Tinker.etal.2005, Pillepich.etal.2008}, leading to a new set of fitting formulae for the mean biasing relation, and a better understanding of the origin of halo biasing. Independently of the exact halo definition, assumed cosmology, simulation box size and resolution, there is a consensus that in a cold-dark matter scenario: i) at a given epoch, more massive haloes are more biased tracers of the underlying matter than lower mass haloes; ii) for halos of fixed mass, the amount of biasing increases with redshift. However, it is still a huge step from a successful description of ``halo biasing'' to that of ``galaxy biasing'' as the latter requires incorporating a recipe for galaxy formation (and evolution) within the current cosmological framework. Galaxy biasing has been studied through hydrodynamic simulations \\citep[e.g.][]{Blanton.etal.1999, Blanton.etal.2000, Cen&Ostriker.2000, Yoshikawa.etal.2001} and semi-analytical modelling combined with N-body simulations \\citep[e.g.][hereafter SBD]{Kauffmann.etal.1997, Benson.etal.2000, Somerville.etal.2001, Sigad.etal.2000}. Despite the difference in the treatment of the various gas-related processes,", "conclusions": "\\label{sec_concl} In this work, we make use of the reconstructed overdensity field in the zCOSMOS volume \\citep[see][]{Kovac.etal.2009} to derive the conditional mean function $\\langle \\delta_g|\\delta \\rangle = b(\\delta, z, R) \\delta$ and the second moments of the mean biasing function $b(\\delta, z, R)$. For this purpose we employ the density field on a grid reconstructed by using the three dimensional distances between galaxies and grid points and counting the objects within a spherical top-hat aperture. We implement a novel method ZADE \\citep{Kovac.etal.2009} to account for galaxies not yet observed spectroscopically in the selected samples of galaxies used to reconstruct the density field. For a biasing analysis, the main advantage of the ZADE method is that in a statistical sense, the mean intergalaxy separation is that of all galaxies in the selected galaxy sample, and not only of the sample of galaxies with spectroscopic redshifts. We have carried out a number of tests on the mock catalogues to assess various errors which are going to affect our biasing analysis. Particularly, we have empirically estimated uncertainties due to cosmic variance, shot noise errors and the density field reconstruction errors. \\begin{itemize} \\item Cosmic variance errors cause a spread in the $\\langle \\delta_g|\\delta \\rangle$ function: for a given $\\delta$ there is a range of $\\delta_g$ values measured in the mock catalogues. Quantifying this spread by the standard deviation $\\sigma$ of $\\log(1 + \\delta_g)$, we find that $\\sigma$ is largest in the most underdense regions where $\\sigma \\sim 0.1$, becomes lower at the intermediate $\\delta$ values, $\\sigma < 0.05$, and increaseas again in the most overdense regions up to $\\sigma \\sim 0.05$ at a given $\\delta$. \\item The shot noise (discrete galaxy sampling) errors modify significantly the shape of the $\\langle \\delta_g|\\delta \\rangle$ function in the most underdense regions, making the local bias $b(\\delta, z, R)$ values in the same regions to appear higher. \\item Reconstruction errors are relevant only in the underdense regions. At the reconstructed value of $\\log(1+\\delta_g)=-1$ in the galaxy density field, the reconstruction error can cause an uncertainty of the order of 0.1 in the matter density field $\\log(1+\\delta)$. \\item The $\\hat{b}$ parameter increases due to the shot noise and reconstruction errors. The $\\tilde{b}/\\hat{b}$ parameter is not susceptible to either of these errors. The cosmic variance causes a spread in the measured values of both of these parameters. \\end{itemize} We can summarise our main findings in the biasing analysis of the 10k zCOSMOS galaxies as follows: \\begin{itemize} \\item The conditional mean function $\\langle \\delta_g|\\delta \\rangle$ has a characteristic shape as described below. In most underdense regions, the mean biasing function vanishes. At some $\\delta < 0$, the mean biasing function appears and then rises sharply in the underdense regions, with the local slope of the biasing function larger than unity. Starting from around mean density and towards higher overdensities, the $\\langle \\delta_g|\\delta \\rangle$ function closely follows a linear relation $\\delta_g = b \\delta$ with $b$ a constant. In the most overdense regions zCOSMOS galaxies are antibiased, i.e. locally $b(\\delta,z,R)<1$. This is true for all samples of tracer galaxies used. The conditional mean function $\\langle \\delta_g|\\delta \\rangle$ is clearly nonlinear in the most overdense and underdense regions. \\item There is a detectable change in the shape of the $\\langle \\delta_g|\\delta \\rangle$ function with redshift in the overdense regions. For a given population, galaxies become more biased tracers of the matter in the regions of mean and mildly positive overdensities as redshift increases from 0.4 to 1. There is an indication of an evolution in the value of $\\delta$ in the overdense regions, at which galaxies become antibiased. For a given population of tracer galaxies, this happens at higher $\\delta$ for higher $z$. Taking into account all the sources of errors, we cannot discriminate if the shape of the $\\langle \\delta_g|\\delta \\rangle$ function in the underdense regions stays constant or undergoes some evolution in $0.40.5 \\rm ~km)\\sim 2.47\\times10^{10}$ deg$^{-2}$ at 95\\% confidence limit. ", "introduction": "\\label{sec:intro} Since the first Kuiper Belt object (not including Pluto) was discovered in 1992 \\citep{jewitt1993}, more than 1000 trans-Neptunian objects (TNOs) have been found\\footnote{See http://www.cfa.harvard.edu/iau/lists/TNOs.html for a list of these objects.}. A picture is emerging of the principal characteristics of the TNOs that is rich in information but leaves many unanswered questions. For example, a number of dynamical groups have been identified: classical, resonant, scattered disk \\citep{Luu1997}, and more recently the extended scattered disk \\citep{gladman2002}. What is the number and size distribution of the smaller objects of these dynamical groups? Is there an extension of the Kuiper Belt beyond 50 AU comprising bodies too small to have been detected in direct surveys? Meanwhile, could a possible close stellar encounter in the early history of the Solar System \\citep{Allen2001, Trujillo2001, bernstein2004, Fuentes2008} be responsible for the mass deficit and the depletion of larger objects (D $>$ 150~km) beyond 50~AU? The size distribution of these small TNOs provides important clues to the dynamical evolution of the early Solar System. Kilometer-size TNOs with 4\\% albedo are expected to have $M_{\\rm R} \\gtrsim30$, which is still way below the detection limit of the largest ground-based telescopes. Yet the vast majority of the TNO population is beyond the limit of direct observation --for example, the Keck pencil beam survey \\citep{Chiang1999} has a detection limit of $M_{\\rm R} \\approx 27.5$. A number of authors \\citep{Bailey1976, Roques2000} have suggested the possibility of detecting TNOs by stellar occultation. An occultation manifests itself as the shadow created by a TNO occulting a background star sweeping across the Earth, causing flux reduction in the lightcurve. Hence, as opposed to direct observation, stellar occultation by TNOs provides a unique way of detecting kilometer to sub-kilometer size objects in the foreseeable future. Recently a few groups have attempted to search for occultations by TNOs \\citep{Roques2006,bickerton2008,Lehner2009,kiwi2008,Bianco2009,bickerton2009}, but no real detection has been confirmed yet. Projects committed to search for TNOs by occultation either do not have adequate time resolution, observe too few stars with good SNR or do not have enough observing time. These projects do a blind search for occultations, and these events are rare due to the reasons mentioned above. Therefore, the key to a successful detection is to have high sampling ($\\ge$~20~Hz), high signal to noise ratio of background stars (SNR $\\ge$ 80), small stellar angular sizes ($\\leq$ 0.1~mas) and, most importantly, many star-hours of observations ($>$ a few hundred thousands). The Panoramic Survey Telescope and Rapid Response System (Pan-STARRS) is a project consisting of 4 telescopes that can cover over 6000 $\\rm deg^{2}$ per night or scan the whole visible sky from Hawaii ($3\\pi$) in a week to a detecting limit $\\sim 24^{\\rm th}$ magnitude. Pan-STARRS prototype telescope (PS1) will monitor up to 60 guide stars in a very high sampling rate video mode. We have compiled a list which allows us to select guide stars from anywhere in the field; guide stars could be in any of the 64$\\times$64 OTCCDs (see $\\S$\\ref{sec:ps1}) and read at $\\sim30$~Hz. Each field will have multiple choices of guide stars that were ranked based on the predicted event rates, which depend on their SNRs and angular sizes. By properly selecting guide stars, the Pan-STARRS video mode images would be ideal for searching TNOs. In the next section, we briefly describe the Pan-STARRS system. In $\\S$\\ref{sec:diff} and $\\S$\\ref{sec:det} we discuss the diffraction profiles and present our detection algorithm. In $\\S$ \\ref{sec:angular} we describe the prediction of stellar angular size using $(V-K)$ color. In $\\S$\\ref{sec:rate} we estimate the number of expected events based on different number density estimations, sampling rates and stellar angular sizes. The methodology and compilation of the guide star list is in $\\S$\\ref{sec:catalog}. In $\\S$\\ref{sec:engdata}, we show and describe the quality of the engineering data obtained in fall 2008. Conclusions are in the final section. ", "conclusions": "\\label{sec:result} We have made a pre-survey study of using lightcurves from guide star video mode images to search for occultation by TNOs near 43~AU. Simulations were made to calculate the null hypothesis distribution (\\fig{fig:nulldist1}) and detection efficiency for various angular sizes and SNRs. Under 30~Hz sampling, PS1 can detect objects as small as $\\sim400$ m at 43~AU (\\fig{fig:efficiency}). Using $(V-K)$ indices, we predicted the angular sizes of all matched stars between Tycho2 and 2MASS catalogs with $m_{\\rm V}<13$ mag and $m_{\\rm K}<16$ mag above the PS1 southern declination limit $\\delta \\ge-30^{\\circ}$. The distribution of stellar angular sizes peak around 0.02~mas in PS1 $3\\pi$ sky (\\fig{fig:vmk}). On average, there are about 180 stars with $m_{\\rm V}<11.5$ mag for PS1 7 deg$^{2}$ field of view. Given that we can choose only 60 guide stars for each PS1 target field, we ranked our guide star candidates by their number of expected events based on their angular sizes, SNRs and model by P\\&S05. As mentioned before, multiple choices of guide star will ensure us the freedom of selecting guide stars away from gaps in the CCD array or dead pixels that would develop over time. Based on the differential size distribution from different models and the threshold to have one false positive in the PS1 three-year lifetime, we estimated the number of expected events could be somewhere from 1 to $\\sim100$ (\\fig{fig:rate1}). The engineering data allowed us to investigate the quality of the lightcurve and develop the detection pipeline for the upcoming real data. We have established that the detection technique performs as well with the filtered engineering data. We also realize that the true SNR is limited by systematics. Some of the systematics will be removed as we move to real data stream, however some of the systematics will remain. For example, scintillation will limit the SNR to about 200. We have recalculated the event rates with the worst case scenario and found that the event rates were compromised by a factor of four. Even with this pessimistic estimation the event rate that PS1 will find can allow us to place a constraint on the size distribution and hence the evolution of the TNOs. Using the available engineering data and detection efficiency at 0.4 and 0.5~km, we were able to derive an effective solid angle $\\sim 2.3\\times 10^{-11}$ and $\\sim 1.2\\times 10^{-10}$ deg$^{2}$ and set the 95\\% confidence upper limit on surface number density at $N(>0.4 \\, \\rm km)\\sim 1.3\\times10^{11}$ and $N(>0.5 \\, \\rm km)\\sim 2.47\\times10^{10}$ deg$^{-2}$ (\\fig{fig:upper}). In future work, for the upcoming real data, we will set the threshold based on maximum true positive to false positive ratio, which allows more candidate events for further investigations. Meanwhile, the video mode lightcurves can also be used to search for objects in Sedna like orbits from 100 to 1000~AU. We will work on a new detection algorithm that is capable of searching for objects in this region." }, "0910/0910.2976_arXiv.txt": { "abstract": "{ Since a century cosmic rays are based on direct cosmic particle detection in space (below PeV) or on secondary downward vertical airshowers (above TeVs). We consider the guaranteed physics of horizontal (hadron) air-showers, HAS, developing at high ($30-40$) km altitudes, above and below these energy windows. Their morphology and information traces are different from vertical ones. Hundreds of km long HAS are often split by geomagnetic fields in a long (fan-like) showering with a twin spiral tail. The horizontal fan-like airshowers are really tangent and horizontal only at North and South poles. At different latitude their showering plane are turned and inclined by geo-magnetic fields. In particular at magnetic equator such tangent horizontal East-West airshowers are bent and developed into a vertical fan air-shower, easily detectable by a vertical array detector (hanging elements by gravity). Such \\emph{medusa } arrays maybe composed by inflated floating balloons chains. The light gas float and it acts as an calorimeter for the particles, while it partially sustains the detector weight. Vertically hanging chains as well as rubber doughnut balloons ( whose interior may record Cherenkov lights) reveal bundles of crossing electron pairs. Even in space orbit such vertical array may hang by tiny tidal forces within huge balloon arrays, while brought in locus by an extendable measure tape. Possibly located around Space Station in synergy with future AMS-particle detector. Such an array maybe loaded at best and cheapest prototype in common balloons tracing upward and tangent hadron air-showers from terrestrial atmosphere edge. These array structure may reveal the split shower signature. Offering a way to disentangle better their shower origination, energy and interaction point. Better revealing the composition nature. Just beyond the Earth horizons there are exciting, but rarer, sources of upward airshowers: the new UHE Tau Air-showers astronomy originated within Earth by EeVs tau neutrino signals skimming the soil and forming UHE Tau, decaying later in flight. Therefore Horizontal airshowers at equator may show the hadron horizontal twin split nature, its composition as well as a very first expected UHE Neutrino astronomy.\\ } \\FullConference{European Physical Society Europhysics Conference on High Energy Physics, EPS-HEP 2009,\\\\ July 16 - 22 2009\\\\ Krakow, Poland} \\begin{document} ", "introduction": " ", "conclusions": "" }, "0910/0910.0468_arXiv.txt": { "abstract": "Giant planet formation process is still not completely understood. The current most accepted paradigm, the core instability model, explains several observed properties of the solar system's giant planets but, to date, has faced difficulties to account for a formation time shorter than the observational estimates of protoplanetary disks' lifetimes, especially for the cases of Uranus and Neptune. In the context of this model, and considering a recently proposed primordial solar system orbital structure, we performed numerical calculations of giant planet formation. Our results show that if accreted planetesimals follow a size distribution in which most of the mass lies in 30-100 meter sized bodies, Jupiter, Saturn, Uranus and Neptune may have formed according to the nucleated instability scenario. The formation of each planet occurs within the time constraints and they end up with core masses in good agreement with present estimations. ", "introduction": " ", "conclusions": "" }, "0910/0910.4282_arXiv.txt": { "abstract": "We present a coherent and homogeneous multi-line study of the CS molecule in nearby (D$<$10Mpc) galaxies. We include, from the literature, all the available observations from the $J=1-0$ to the $J=7-6$ transitions towards NGC~253, NGC~1068, IC~342, Henize~2-10, M~82, the Antennae Galaxies and M~83. We have, for the first time, detected the CS(7-6) line in NGC~253, M~82 (both in the North-East and South-West molecular lobes), NGC 4038, M~83 and tentatively in NGC~1068, IC~342 and Henize~2-10. We use the CS molecule as a tracer of the densest gas component of the ISM in extragalactic star-forming regions, following previous theoretical and observational studies by Bayet et al. (2008a,b and 2009). In this first paper out of a series, we analyze the CS data sample under both Local Thermodynamical Equilibrium (LTE) and non-LTE (Large Velocity Gradient-LVG) approximations. We show that except for M~83 and Overlap (a shifted gas-rich position from the nucleus NGC~4039 in the Antennae Galaxies), the observations in NGC~253, IC~342, M~82-NE, M~82-SW and NGC~4038 are not well reproduced by a single set of gas component properties and that, at least, two gas components are required. For each gas component, we provide estimates of the corresponding kinetic temperature, total CS column density and gas density. ", "introduction": "\\label{sec:intro} Detection of star-forming gas is one of the most direct ways to measure the star formation rate and activity in a galaxy, allowing us to significantly improve our understanding of galaxy formation and evolution. Star-forming regions of very dense gas (n(H$_{2}$)$> 10^{5}$cm$^{-3}$) are needed to maintain star formation activity, even in hostile environments associated with young massive stars. In these environments, star-forming regions are able to resist the disruptive forces (winds or radiation) from nearby newly formed stars longer than the local gas in the local interstellar medium \\citep{Klei83, LaRo83}. Determining the physical conditions of the very dense gas in which massive stars form is, thus, essential. In this paper, we study such dense gas and estimate its properties over a large range of nearby galaxy types. Following theoretical studies by \\citet{Baye08a} and the first detections of extragalactic very dense gas presented in \\citet{Maue89a, Maue89b, Walk90, Mart05, Mart06b, Mart06a, Baye08b, Grev09}, we have carried out multi-line observations of the CS molecule in nearby (D$<$10Mpc) extragalactic environments, enhancing significatively the current data set of extragalactic CS observations. Indeed, CS lines have not been observed so far but in few brightest nearby nuclei such as NGC~253, IC~342 and M~82 and mainly in their lower-J rotational levels (see references above). Hence, the physical properties of such galaxies namely the kinetic temperature, gas density, etc were estimated using only a small sample of CS lines. Gas traced by higher-J of species such as CS has not be characterized so far. In addition, even for the brightest sources (e.g. NGC~253 and M~82), no study of the various velocity components nor various positions in the same galaxy has been performed so far in a systematic way. In this paper, first of a series, we thus aim at investigating in much more details the very dense gas properties in many extragalactic sources as traced by the CS molecule. Sulphur-bearing species are shown to be particularly enhanced during massive star formation, while species such as HCN, although a useful dense gas tracer, may not be tracing the sites where star formation occurs \\citep{Lint06}. Recently, \\cite{Mart05} showed that sulfur emission in the nuclear region of the nearby starburst galaxy NGC~253 is very strong. Amongst sulfur-bearing species, the CS molecule appears as one of the best tracers of dense gas, especially its high-J rotational transitions such as the CS $J=5-4$ line with an excitation threshold higher than 10$^{4}$-10$^{5}$ cm$^{-3}$ \\citep{Bron96} and the CS $J=7-6$ line with a critical density of n$_{crit} \\sim 2\\times 10^{7}$ cm$^{-3}$ \\citep{Plum92}. Selection of nearby sources where we have proposed to observe the CS lines was made through a comparison with past and recent CS observations (see references above). Sources where no CS detections were available from the literature (e.g. Henize 2-10) were chosen on the basis of the detections of the CO(6-5) or CO(7-6) lines since these lines are also good tracers of a relatively dense and quite warm gas \\citep{Baye04, Baye06}. The source selection also aimed at presenting various nearby galaxy types where CS line emission can be studied and compared. This sample is far from being unbiased but it gathers so far the most complete CS data sample ever obtained in extragalactic sources. Thus, the centers of NGC~253, IC~342, Henize~2-10, the two molecular lobes of M~82 (North-East and South-West positions), three positions in the Antennae Galaxies (the two nuclei: NGC~4038 and NGC~4039, and a shifted position from NGC~4039 called Overlap) and the center of M~83 were selected, covering a large range of galaxy types and star-formation activities (starburst, irregular, merging galaxies, more quiet star-forming galaxies, etc). We have also included in our source sample the center of NGC~1068 (AGN-dominated galaxy) where HCN has been detected \\citep{Plan91,Krip08}. We present our observations, data reduction and results in Sect. \\ref{sec:obs}. In Sect. \\ref{sec:mol} we analyze the data set both under Local Thermodynamical Equilibrium (LTE) and non-LTE (Large Velocity Gradient-LVG) approximations (see Subsects. \\ref{subsec:lte} and \\ref{subsec:lvg}). In Sects. \\ref{sec:disc} and \\ref{sec:con} we discuss our findings and conclude, respectively. ", "conclusions": "\\label{sec:con} In order to better determine the properties of the very dense star-forming gas over a large range of physical conditions, we present the most complete, extragalactic CS survey ever performed on nearby galaxies. In particular, we detect for the first time the CS(5-4) and CS(7-6) lines, in various environments such as starburst, AGN-dominated galaxies, irregular galaxies, merging galaxies, etc. In this first paper out of a series, we simply analyzed the data through a rotational diagram method and an LVG modelling and we show that for most sources at least two gas components are needed to reproduce the observations. The low-temperature gas component properties, derived from the CS(2-1), CS(3-2) and CS(4-3) lines (low-J CS lines), vary from source to source within the range T$_{K}$, n(H$_{2}$), N(CS) = 10-30K, 0.16-1.60$\\times 10^{5}$cm$^{-3}$, 0.25-6.30$\\times 10^{14}$cm$^{-2}$, respectively. The high-temperature gas component properties, derived from the CS(5-4) and CS(7-6) lines (high-J CS lines), vary from source to source within the range T$_{K}$, n(H$_{2}$), N(CS) = 45-70K, 6.30-40.0$\\times 10^{5}$cm$^{-3}$, 0.16-1.00$\\times 10^{14}$cm$^{-2}$, respectively. We also show that, using the current data set of CS observations, in Overlap and M~83, it appears that the very dense gas may be more homogeneously distributed than in other nearby sources.However, we again underline here that our estimates of the physical properties of these galaxies suffer from LVG degeneracy. Comparison with other tracers of dense gas such as HCN, HNC, HCO$^{+}$, etc even in the brightest sources of our sample is very difficult to perform. However, in the molecular lobes of M~82, we have been able to show a good correlation between the CS and the methanol. In both NGC~253-1 and NGC~253-2, we showed that the HCN and the low-temperature CS gas on one side, and the HCO$^{+}$, H$_{2}$CO, HNC and the high-temperature CS gas on the other side have compatible rotational temperatures and column densities. In pure AGN-dominated galaxy such as NGC~1068, we find no correlation between the HCN and CS. Forthcoming papers will focuss on detailed studies of the very dense gas properties in individual nearby sources, investigating for instance their dense gas star-formation efficiencies, their dense gas masses based on CS modelling, etc, as well as on a more statistical analysis such as Kennicutt-Schmidt laws in both extragalactic and galactic sources. \\begin{landscape} \\begin{table} \\caption{Observational parameters and Gaussian fits for the data set of CS observations. Where data were not available, we have put a black dash in the corresponding place in the table.}\\label{tab:obs} \\resizebox{18.8cm}{!}{ \\begin{tabular}{l c c c c c c c c c c c c c} \\hline Source & Line & $\\nu$ & Tsys & beam & Telesc. & $\\int$(T$_{mb}$ dv) & V$_{peak}$ & FWHM & T$_{peak}$ & rms & RA(J2000) & DEC(J2000) & Refs$^{1}$\\\\ & & & & size & name & & & & & & & & \\\\ & & (GHz) & (K) & ($''$) & & (Kkms$^{-1}$) & (kms$^{-1}$) & (kms$^{-1}$) & (mK) & (mK) & (h:m:s) & ($^{\\circ}$ : ' : '') &\\\\ \\hline NGC 253-1 & CS(2-1) & 97.980 & - & 51.0 & SEST & 4.4$\\pm$0.5 & 172 & 96 & 43.4 & 5.2 & 00:47:33.4 & -25:17:23.0 & d\\\\ & CS(3-2) & 146.969 & - & 16.7 & IRAM-30m & 11.9$\\pm$0.2 & 185 & 100 & 111.2 & 4.1 & '' & '' & d\\\\ & CS(4-3) & 195.954 & - & 12.6 & IRAM-30m & 11.5$\\pm$0.4 & 185 & 100 & 108.2 & 7.0 & '' & '' & d\\\\ & CS(5-4) & 244.936 & - & 21.0 & SEST & 4.4$\\pm$0.4 & 158 & 107 & 38.2 & 10.8 & '' & '' & d\\\\ & CS(7-6) & 342.883 & 679 & 14.0 & JCMT & 7.9$\\pm$0.3 & 185$^{2}$ & 110.1$\\pm$4.4 & 67.1 & 5.1 & '' & '' & a\\\\ \\hline NGC 253-2 & CS(2-1) & 97.980 & - & 51.0 & SEST & 8.2$\\pm$0.5 & 290 & 116 & 66.5 & 5.2 & 00:47:33.4 & -25:17:23.0 & d\\\\ & CS(3-2) & 146.969 & - & 16.7 & IRAM-30m & 13.7$\\pm$0.2 & 288 & 117 & 110.3 & 4.1 & '' & '' & d\\\\ & CS(4-3) & 195.954 & - & 12.6 & IRAM-30m & 13.4$\\pm$0.5 & 288 & 121 & 104.5 & 7.0 & '' & '' & d\\\\ & CS(5-4) & 244.936 & - & 21.0 & SEST & 5.4$\\pm$1.3 & 262 & 108 & 46.7 & 10.8 & '' & '' & d\\\\ & CS(7-6) & 342.883 & 679 & 14.0 & JCMT & 4.4$\\pm$0.3 & 288$^{2}$ & 107.1$\\pm$7.1 & 38.7 & 5.1 & '' & '' & a\\\\ \\hline NGC 1068 & CS(3-2) & 146.969 & 600 & 16.7 & IRAM-30m & 9.1$\\pm$1.5 & 1100 & 245.0 & 30.0 & 10.0 & 02:42:40.7 & -00:00:47.6 & e\\\\ & CS(5-4) & 244.936 & - & 10.0 & IRAM-30m & 3.3$\\pm$0.3 & 1125$\\pm$9.0 & 180.0$\\pm$20.0 & 17.6 & 3.0 & '' & '' & f\\\\ & CS(7-6) & 342.883 & 307 & 14.0 & JCMT & 1.4$\\pm$0.5 & 1125.4$\\pm$49.0 & 279.0$\\pm$99.3 & 4.8 & 8.1 & '' & '' & a\\\\ \\hline IC 342 & CS(1-0) & 48.991 & 350-550 & 36.0 & NRO & 2.5$\\pm$0.4 & 30 & 50-60 & 50.0 & 12.0 & 03:46:48.3 & 68:05:46.0 & g\\\\ & CS(2-1) & 97.980 & 188 & 25.1 & IRAM-30m & 5.0$\\pm$0.1 & 31.7$\\pm$0.6 & 53.9$\\pm$1.5 & 87.4 & 2.5 & '' & '' & c\\\\ & CS(3-2) & 146.969 & 225 & 16.7 & IRAM-30m & 4.9$\\pm$0.1 & 31.2$\\pm$0.6 & 51.7$\\pm$1.6 & 88.6 & 3.0 & '' & '' & c\\\\ & CS(5-4) & 244.936 & 378 & 10.0 & IRAM-30m & 2.5$\\pm$0.2 & 38.2$\\pm$2.0 & 63.4$\\pm$5.1 & 37.6 & 2.9 & '' & '' & c\\\\ & CS(7-6) & 342.883 & 525 & 14.0 & JCMT & 1.0$\\pm$0.4 & 29.3$\\pm$15.3 & 75.6$\\pm$30.6 & 12.3 & 8.5 & '' & '' & a\\\\ \\hline Henize 2-10 & CS(7-6) & 342.883 & 252 & 14.0 & JCMT & 0.6$\\pm$0.2 & 883.1$\\pm$7.4 & 42.0$\\pm$16.1 & 13.0 & 7.4 & 08:36:15.2 & -26:24:34.0 & a\\\\ \\hline M~82\\_NE & CS(2-1) & 97.980 & 236 & 25.1 & IRAM-30m & 9.4$\\pm$0.2 & 299.9$\\pm$1.3 & 103.4$\\pm$3.3 & 85.7 & 8.5 & 09:55:54.4 & 69:40:54.6 & a\\\\ & CS(3-2) & 146.969 & 278 & 16.7 & IRAM-30m & 8.9$\\pm$0.1 & 290.9$\\pm$0.7 & 114.6$\\pm$2.0 & 73.3 & 1.7 & '' & '' & a\\\\ & CS(4-3) & 195.954 & 1425 & 12.6 & IRAM-30m & 6.5$\\pm$0.8 & 309.2$\\pm$6.1 & 87.7$\\pm$12.1 & 70.1 & 17.2 & '' & '' & a\\\\ & CS(5-4) & 244.936 & 772 & 10.0 & IRAM-30m & 6.4$\\pm$0.2 & 313.0$\\pm$1.1 & 73.5$\\pm$2.6 & 82.1 & 4.8 & '' & '' & a\\\\ & CS(7-6) & 342.883 & 309 & 14.0 & JCMT & 1.1$\\pm$0.2 & 313.2$\\pm$4.4 & 58.1$\\pm$6.8 & 15.7 & 2.9 & '' & '' & a\\\\ \\hline M~82\\_SW & CS(2-1) & 97.980 & 237 & 25.1 & IRAM-30m & 8.9$\\pm$0.2 & 131.4$\\pm$1.2 & 109.8$\\pm$2.7 & 75.9 & 3.8 & 09:55:49.4 & 69:40:39.6 & a\\\\ & CS(4-3) & 195.954 & 1628 & 12.6 & IRAM-30m & 7.3$\\pm$1.0 & 140.0$\\pm$6.6 & 104.7$\\pm$17.2 & 65.5 & 17.9 & '' & '' & a\\\\ & CS(7-6) & 342.883 & 340 & 14.0 & JCMT & 1.0$\\pm$0.2 & 110.2$\\pm$8.8 & 60.3$\\pm$14.8 & 13.8 & 5.1 & '' & '' & a\\\\ \\hline NGC 4038 & CS(2-1) & 97.980 & 201 & 25.1 & IRAM-30m & 1.0$\\pm$0.1 & 1646.8$\\pm$3.0 & 79.6$\\pm$6.7 & 11.6 & 2.2 & 12:01:52.8 & -18:52:05.3 & a\\\\ & CS(3-2) & 146.969 & 471 & 16.7 & IRAM-30m & 0.8$\\pm$0.1 & 1645.2$\\pm$3.8 & 62.2$\\pm$7.6 & 11.4 & 2.6 & '' & '' & a\\\\ & CS(4-3) & 195.954 & 1396 & 12.6 & IRAM-30m & 1.1$\\pm$0.3 & 1646.0$\\pm$8.6 & 65.4$\\pm$15.8 & 15.3 & 6.5 & '' & '' & a\\\\ & CS(5-4) & 244.936 & 326 & 20.0 & JCMT & 1.6$\\pm$0.2 & 1649.9$\\pm$6.5 & 103.2$\\pm$12.7 & 14.4 & 6.4 & '' & '' & b\\\\ & CS(5-4) & 244.936 & 1924 & 10.5 & IRAM-30m & 1.9$\\pm$0.4 & 1666.3$\\pm$9.7 & 103.0 & 17.1 & 9.5 & '' & '' & a\\\\ & CS(7-6) & 342.883 & 250 & 14.0 & JCMT & 0.7$\\pm$0.1 & 1629.4$\\pm$7.6 & 88.6$\\pm$15.0 & 7.8 & 2.7 & '' & '' & a\\\\ \\hline NGC 4039 & CS(2-1) & 97.980 & 226 & 25.1 & IRAM-30m & 1.7$\\pm$0.2 & 1640.8$\\pm$16.3 & 228.2$\\pm$33.6 & 7.1 & 3.2 & 12:01:53.5 & -18:53:11.3 & a\\\\ & CS(3-2) & 146.969 & 547 & 16.7 & IRAM-30m & 1.3$\\pm$0.3 & 1691.6$\\pm$19.1 & 145.0$\\pm$30.2 & 8.3 & 5.8 & '' & '' & a\\\\ & CS(4-3) & 195.954 & 1275 & 12.6 & IRAM-30m & - & - & - & - & 17.9 & '' & '' & a\\\\ & CS(5-4) & 244.936 & 277 & 20.0 & JCMT & 1.4$\\pm$0.2 & 1757.6$\\pm$12.7 & 157.1$\\pm$26.5 & 8.2 & 4.0 & '' & '' & a\\\\ \\hline OVERLAP & CS(2-1) & 97.980 & 203 & 25.1 & IRAM-30m & 0.9$\\pm$0.2 & 1498.8$\\pm$5.6 & 81.2$\\pm$18.9 & 10.6 & 2.7 & 12:01:54.9 & -18:52:59.0 & a\\\\ & CS(3-2) & 146.969 & 467 & 16.7 & IRAM-30m & 0.8$\\pm$0.2 & 1504.8$\\pm$6.4 & 61.8$\\pm$18.7 & 11.7 & 4.8 & '' & '' & a\\\\ & CS(4-3) & 195.954 & 1290 & 12.6 & IRAM-30m & - & - & - & - & 14.0 & '' & '' & a\\\\ & CS(5-4) & 244.936 & 1763 & 10.0 & IRAM-30m & - & - & - & - & 14.4 & '' & '' & a\\\\ & CS(7-6) & 342.883 & 230 & 14.0 & JCMT & - & - & - & - & 4.3 & '' & '' & a\\\\ \\hline M~83 & CS(3-2) & 146.969 & 600 & 16.7 & IRAM-30m & 1.2$\\pm$0.3 & 553.0$\\pm$11.0 & 78.0$\\pm$24.0 & 15.0 & - & 13:36:59.2 & -29:52:04.5 & e\\\\ & CS(5-4) & 244.936 & - & 10.0 & IRAM-30m & 2.3$\\pm$0.3 & 504.0$\\pm$7.0 & 82.4$\\pm$1.1 & 26.5 & 0.5 & '' & '' & d\\\\ & CS(7-6) & 342.883 & 121 & 14.0 & JCMT & 0.4$\\pm$0.1 & 540.0$\\pm$9.5 & 99.1$\\pm$18.8 & 3.3 & 1.1 & '' & '' & a\\\\ \\hline \\end{tabular}} $^{1}$ Refs: a: This work, b: See \\citet{Baye08b}, c: See \\citet{Alad09}, d: \\citet{Mart05}, e: \\citet{Maue89b}, f: \\citet{Mart09} and g: \\citet{Pagl95}; $^{2}$: The velocity positions of the two components in NGC 253 have been fixed to 185 kms$^{-1}$ and 288 kms$^{-1}$, consistently to what it has been performed in \\citet{Mart05}. \\end{table} \\end{landscape} \\begin{table*} \\caption{Results of the linear regressions (slope, correlation coefficient, total source-averaged CS column density and rotational temperature) for a single- and two-components fit. The linear regressions have been obtained using the xmgrace software, including in the calculations the error bars of the observations.}\\label{tab:T_rot} \\begin{tabular}{ccccc} \\hline Source (Nb. of & slope & Correl. & N(CS)$^{1}$ & T$_{rot}$ $^{1}$\\\\ fit components) & ($\\times 10^{-2}$) & Coeff. & ($\\times 10^{14}$) & (in K)\\\\ \\hline NGC 253-1 (one) & -2.67 & 0.878 & 1.79 & 16.3\\\\ NGC 253-1 (two) & -7.14 & 0.970 & 2.65 & 6.1 \\\\ & -1.33 & 0.969 & 0.80 & 32.7 \\\\ \\hline NGC 253-2 (one) & -3.40 & 0.910 & 2.47 & 12.8\\\\ NGC 253-2 (two) & -8.34 & 0.952 & 4.39 & 5.2\\\\ & -2.14 & 0.999 & 0.97 & 20.3 \\\\ \\hline NGC 1068 (one) & -3.13 & 0.921 & 2.88 & 13.9\\\\ NGC 1068 (two) & -6.09 & 1 & 5.80 & 7.1\\\\ & -1.31 & 1 & 0.68 & 33.2\\\\ \\hline IC 342 (one) & -4.26 & 0.936 & 2.67 & 10.2\\\\ IC 342 (two) & -10.87 & 0.997 & 2.87 & 4.0 \\\\ & -1.50 & 1 & 0.27 & 28.9 \\\\ \\hline M82-NE (one) & -3.91 & 0.958 & 2.64 & 11.1\\\\ M82-NE (two) & -7.81 & 0.992 & 4.25 & 5.6\\\\ & -2.77 & 0.999 & 1.04 & 15.7\\\\ \\hline M82-SW (one) & -3.94 & 0.975 & 1.80 & 11.0\\\\ M82-SW (two) & -7.10 & 1 & 2.39 & 6.1\\\\ & -3.02 & 1 & 0.74 & 14.4\\\\ \\hline size=7 $''$ & & & & \\\\ NGC 4038 (one) & -2.38 & 0.895 & 0.25 & 18.3 \\\\ NGC 4038 (two) & -6.24 & 0.936 & 0.29 & 7.0\\\\ & -1.52 & 0.991 & 0.15 & 28.6\\\\ size=25\\%$\\times$7 $''$ & & & & \\\\ NGC 4038 (one) & -2.43 & 0.891 & 0.42 & 17.9 \\\\ NGC 4038 (two) & -6.49 & 0.943 & 0.51 & 6.7\\\\ & -1.48 & 0.993 & 0.24 & 29.2\\\\ size=50\\%$\\times$7 $''$ & & & & \\\\ NGC 4038 (one) & -2.46 & 0.887 & 0.88 & 17.6 \\\\ NGC 4038 (two) & -6.67 & 0.947 & 1.10 & 6.5\\\\ & -1.45 & 0.995 & 0.49 & 29.9\\\\ size=75\\%$\\times$7 $''$ & & & & \\\\ NGC 4038 (one) & -2.47 & 0.885 & 0.34 & 17.6 \\\\ NGC 4038 (two) & -6.74 & 0.949 & 4.29 & 6.5\\\\ & -1.42 & 0.997 & 1.85 & 30.5\\\\ variable size & & & & \\\\ NGC 4038 (one) & -1.90 & 0.885 & 0.36 & 22.9 \\\\ NGC 4038 (two) & -5.40 & 0.962 & 0.34 & 8.0\\\\ & -1.16 & 0.928 & 0.26 & 37.5\\\\ \\hline NGC 4039 (one) & -3.21 & 0.847 & 0.81 & 13.5 \\\\ NGC 4039 (two) & -11.34 & 1 & 1.01 & 3.8 \\\\ & -1.27 & 1 & 0.55 & 34.1 \\\\ \\hline Overlap (one) & -5.67 & 1 & 0.07 & 7.7\\\\ \\hline M 83 (one) & -2.73 & 0.999 & 0.45 & 15.9\\\\ \\hline \\end{tabular} $^{1}$: The total source-averaged CS column density is derived using the following formulae: N(CS)=$10^{intersect} \\times$Q$_{rot}$(T$_{rot}$) where \\textit{intersect} is the value of the intersection of the linear regression with the y-axis (see Figs. \\ref{fig:11} to \\ref{fig:16}) and Q$_{rot}$(T$_{rot}$) is the partition function at the rotational temperature T$_{rot}$ (linear interpolation realized when the available range of T$_{rot}$ was not relevant for the studied case - see text). The rotational temperature is calculated using the formulae: T$_{rot} = - \\frac{1}{slope} \\times log(e)$. \\end{table*} \\begin{table} \\caption{Results of the LVG model analysis: the physical properties of the best LVG model (having the lowest-$\\chi^{2}$ value) for both the low- and high-temperature gas components. We remind the reader that these values are only indicative since degeneracy in LVG models occurs.}\\label{tab:LVG} \\begin{center} \\begin{tabular}{ccccc} \\hline Source & T$_{K}$ & n(H$_{2}$) & N(CS) \\\\ & in K & ($\\times 10^{5}$ cm$^{-3}$) & ($\\times 10^{14}$cm$^{-2}$) &\\\\ \\hline NGC 253-1 & 30 & 0.40 & 2.50 \\\\ & 65 & 25.0 & 0.63 \\\\ \\hline NGC 253-2 & 15 & 0.16 & 4.00\\\\ & 70 & 10.0 & 1.00\\\\ \\hline NGC 1068 & 20 & 1.50 & 6.30\\\\ & 65 & 40.0 & 0.63\\\\ \\hline IC 342 & 15 & 0.63 & 1.00 \\\\ & 50 & 16.0 & 0.25 \\\\ \\hline M82-NE & 10 & 1.00 & 4.00 \\\\ & 65 & 6.30 & 1.00 \\\\ \\hline M82-SW & 15 & 1.60 & 2.50 \\\\ & 45 & 6.30 & 0.63\\\\ \\hline NGC 4038 & 10 & 1.00 & 0.25 \\\\ & 50 & 40.0 & 0.16 \\\\ \\hline NGC 4039 & 10 & 0.40 & 1.00 \\\\ & 45 & 2.50 & 0.40 \\\\ \\hline Overlap & 10 & 2.50 & 0.06 \\\\ \\hline M 83 & 65 & 6.30 & 0.15 \\\\ \\hline \\end{tabular} \\end{center} \\end{table} \\begin{figure} \\includegraphics[height=3cm]{fig1.jpg} \\caption{CS(7-6) spectrum in the center of NGC~253. The angular resolution is shown in the upper right corner. The thick vertical black line symbolizes the V$_{LSR}$ of the source. For NGC~253, we have applied a two-components fit (thin black lines) after having smoothed the entire signal to a velocity resolution of $\\approx 10$ kms$^{-1}$. The fit, on the left is for the emission from NGC~253-1 while on the right, it is for the NGC~253-2 emission.}\\label{fig:1} \\end{figure} \\begin{figure} \\includegraphics[height=3cm]{fig2.jpg} \\caption{Marginal detection of the CS(7-6) line in the center of NGC~1068 ($\\Delta$v$=13.7$kms$^{-1}$).}\\label{fig:2} \\end{figure} \\begin{figure} \\includegraphics[height=3cm]{fig3.jpg} \\caption{Marginal detection of the CS(7-6) line in the center of IC~342 ($\\Delta$v$=13.7$kms$^{-1}$).}\\label{fig:3} \\end{figure} \\begin{figure} \\includegraphics[height=3cm]{fig4.jpg} \\caption{Marginal detection of the CS(7-6) line in the center of Henize~2-10 ($\\Delta$v$=13.7$kms$^{-1}$).}\\label{fig:4} \\end{figure} \\begin{figure} \\includegraphics[height=14cm]{fig5.jpg} \\caption{CS spectra in the NE molecular lobe of M~82. The velocity resolution from top to bottom is $\\Delta$v$=12.2$kms$^{-1}$, 8.1kms$^{-1}$, 24.5kms$^{-1}$, 9.7kms$^{-1}$ and 13.7kms$^{-1}$, respectively.}\\label{fig:5} \\end{figure} \\begin{figure} \\includegraphics[height=9cm]{fig6.jpg} \\caption{CS spectra in the SW molecular lobe of M~82. The velocity resolution from top to bottom is $\\Delta$v$=12.2$kms$^{-1}$, 24.5kms$^{-1}$ and 13.7kms$^{-1}$, respectively.}\\label{fig:6} \\end{figure} \\begin{figure} \\includegraphics[height=18cm]{fig7.jpg} \\caption{CS spectra in NGC~4038 (one of the nuclei in The Antennae Galaxies). The velocity resolution from top to bottom is $\\Delta$v$=24.5$kms$^{-1}$, 16.3kms$^{-1}$, 24.5kms$^{-1}$, 10.0kms$^{-1}$, 10.0kms$^{-1}$ and 13.7 kms$^{-1}$, respectively.}\\label{fig:7} \\end{figure} \\begin{figure} \\includegraphics[height=12cm]{fig8.jpg} \\caption{CS spectra in NGC~4039 (one of the nuclei in The Antennae Galaxies). The velocity resolution from top to bottom is $\\Delta$v$=24.5$kms$^{-1}$, 16.3kms$^{-1}$, 12.2kms$^{-1}$ and 19.1kms$^{-1}$, respectively.}\\label{fig:8} \\end{figure} \\begin{figure} \\includegraphics[height=15cm]{fig9.jpg} \\caption{CS spectra in the Overlap position (The Antennae Galaxies). Due to poor weather conditions, no detection is seen neither in the CS(4-3), nor the CS(5-4) nor the CS(7-6) lines. The velocity resolution from top to bottom is $\\Delta$v$=24.5$kms$^{-1}$, 16.3kms$^{-1}$, 24.5kms$^{-1}$, 19.6kms$^{-1}$ and 13.7kms$^{-1}$, respectively.}\\label{fig:9} \\end{figure} \\begin{figure} \\includegraphics[height=3cm]{fig10.jpg} \\caption{CS(7-6) spectrum in the center of M~83 ($\\Delta$v$=13.7$kms$^{-1}$).}\\label{fig:10} \\end{figure} \\begin{figure} \\includegraphics[height=10.5cm]{fig11.jpg} \\caption{CS rotational diagrams of NGC~253-1 (top plot) and NGC~253-2 (bottom plot). The linear regression using a single-component is represented by dashed black lines while the two-components regression is represented by solid black lines. Detections are represented by black squares symbols with error bars. These error bars actually correspond to those of the integrated intensities (order of 10-20\\%). When a CS line is marginally detected (upper limit), we represent the corresponding datum with an open white square (see also Figs. \\ref{fig:12}, \\ref{fig:13} and \\ref{fig:15}).}\\label{fig:11} \\end{figure} \\begin{figure} \\includegraphics[height=5.5cm]{fig12.jpg} \\caption{CS rotational diagram of the center of NGC~1068.}\\label{fig:12} \\end{figure} \\begin{figure} \\includegraphics[height=5.5cm]{fig13.jpg} \\caption{CS rotational diagram of the center of IC~342.}\\label{fig:13} \\end{figure} \\begin{figure} \\includegraphics[height=10cm]{fig14.jpg} \\caption{CS rotational diagrams for the M~82-NE (top plot) and the M~82-SW (bottom plot) molecular lobes. }\\label{fig:14} \\end{figure} \\begin{figure} \\includegraphics[height=14cm]{fig15.jpg} \\caption{CS rotational diagrams of the three positions observed in the Antennae: NGC~4038 (top plot), NGC~4039 (middle plot) and Overlap (bottom plot). For the NGC~4038 CS(5-4) line, we have used a white open triangle and a white open square for representing the IRAM-30m and the JCMT observations, respectively.}\\label{fig:15} \\end{figure} \\begin{figure} \\includegraphics[height=5cm]{fig16.jpg} \\caption{CS rotational diagram of the center of M~83. }\\label{fig:16} \\end{figure}" }, "0910/0910.3694_arXiv.txt": { "abstract": "% ", "introduction": " ", "conclusions": "" }, "0910/0910.1144_arXiv.txt": { "abstract": "We study the hadron-quark phase transition with the finite size effects at finite temperature. For the hadron phase, we adopt a realistic equation of state in the framework of the Brueckner-Hartree-Fock theory including hyperons. The properties of the mixed phase are clarified by considering the finite size effects under the Gibbs conditions. We find that the equation of state becomes softer than that at zero-temperature for some density region. We also find that the equation of state gets closer to that given by the Maxwell construction. Moreover, the number of hyperons is suppressed by the presence of quarks. These are characteristic features of the hadron-quark mixed phase, and should be important for many astrophysical phenomena such as mergers of binary neutron stars. ", "introduction": "Introduction} Nowadays the effects of quark matter on various astrophysical phenomena have been studied extensively. For example, the cooling of compact stars have been studied in Ref.~\\cite{page00, blaschke00, blaschke01, grigorian05}. Other examples include the effects of quark matter on gravitational wave radiation~\\cite{lin06, yas07, abdikamalov08}, neutrino emissions~\\cite{hatsuda87, nakazato08, sagert08}, rotational frequencies~\\cite{burgio03}, and the energy release during the collapse from neutron stars to quark stars~\\cite{yas05, zdunik07}, etc.. In particular, the mechanisms of supernovae and gamma-ray bursts have not been clearly understood; the QCD phase transition may take place in such phenomena and take critical role~\\cite{gentile93}. However, there are left many uncertainties for the hadron-quark phase transition, e.g. the equation of state (EOS) of quark matter or deconfinement mechanism. Assuming the quark deconfinement transition to be of first order, we find it causes a thermodynamical instability and the mixed phase appears around the critical density. Since there are two conserved quantities, baryon number and electric charge, the phase equilibrium in the mixed phase must be carefully treated by applying the Gibbs conditions \\cite{glendenning92}, instead of the Maxwell construction. A simple treatment of the mixed phase may be the bulk Gibbs calculation, where phase equilibrium of two bulk matter is considered without electromagnetic interaction and surface tension. Generally the properties of the mixed phase should strongly depend on electromagnetic interaction and surface tension, and these effects, sometimes called ``{\\it the finite-size effects}\", lead to the non-uniform \"Pasta\" structures. The EOS of the mixed phase becomes similar to the one under the bulk Gibbs calculation for weak surface tension, and to the one given by the Maxwell construction for strong surface tension~\\cite{voskresensky03,endo06,maruyama07}. The charge screening is also important for their mechanical instability. In the previous papers these finite-size effects have been properly taken into account to elucidate the properties of the pasta structure and demonstrate the importance of the charge screening at zero temperature \\cite{maruyama07}. However, finite temperature comes in many cases such as relativistic heavy-ion collisions and astrophysical phenomena. In this paper, we study the hadron-quark mixed phase with the finite-size effects at finite temperature by extending the previous works of Maruyama et al.~\\cite{maruyama07}. We adopt the Brueckner-Hartree-Fock EOS by Baldo et al. for the hadron phase~\\cite{baldo98}. The EOS includes hyperons as well as nucleons, interacting with hadronic two-body forces and nucleonic three-body forces. For the quark phase, we adopt the thermodynamic bag model for simplicity. We impose the Gibbs conditions on the phase equilibrium, taking into account the finite-size effects. This paper is organized as follows. In Sec.~II, we outline our framework. In Sec.~III, we present numerical results. Sec.~IV is devoted to the conclusion and discussion where we give some astrophysical implications of our results. ", "conclusions": "We have studied the hadron-quark phase transition at finite temperature. We have taken into account the {\\it finite-size effects} imposing the Gibbs conditions on the phase equilibrium, and calculating the density profiles in a self-consistent manner. At finite temperature, EOS of the hadron-quark phase transition gets close to that given by the Maxwell construction. It is due to the mechanical instability of the geometrical structure induced by the thermal effect. Pressure of the mixed phase at finite temperature are 10--30 \\% smaller than that at zero-temperature though the similar behavior appears without hyperons~\\cite{burgio08, yasutake09}. This behavior is characteristic of the hadron-quark mixed phase which can be explained by the large degree of freedom in the quark phase. Hyperon fractions are suppressed by the appearance of the mixed phase, as in the case of zero temperature. Our calculations are subject to the neutrino-free~(low $Y_l$) case at finite temperature. Such situation will appear in mergers of neutron star-neutron star binaries or black hole-neutron star binaries~\\cite{shibata06}, and our result may change their dynamical aspects. Of course, we should take into account isentropic and $Y_l\\neq 0$ situation in the core of supernovae. This work is now in progress. Finally we note again that EOS has many uncertainties, especially for quark matter. We simply adopted the thermodynamic bag model in this paper, and used the density-temperature independent bag constant and surface tension, while it would be interesting to include such dependence for a realistic description~\\cite{baldo03, voskresensky09}. Moreover, chiral restoration or color super conductivities may also change our results~\\cite{kashiwa07,kashiwa07b,jorge07,fukushima08,yasutake09}. These are open questions for astrophysics and nuclear physics." }, "0910/0910.1002_arXiv.txt": { "abstract": "I show how to reintroduce velocity dispersion into perturbation theory (PT) calculations of structure in the Universe, i.e., how to go beyond the pressureless fluid approximation, starting from first principles. This addresses a possible deficiency in uses of PT to compute clustering on the weakly non-linear scales that will be critical for probing dark energy. Specifically, I show how to derive a non-negligible value for the (initially tiny) velocity dispersion of dark matter particles, $\\left<\\delta v^2\\right>$, where $\\delta v$ is the deviation of particle velocities from the local bulk flow. The calculation is essentially a renormalization of the homogeneous (zero order) dispersion by fluctuations 1st order in the initial power spectrum. For power law power spectra with $n>-3$, the small-scale fluctuations diverge and significant dispersion can be generated from an arbitrarily small starting value -- the dispersion level is set by an equilibrium between fluctuations generating more dispersion and dispersion suppressing fluctuations. For an $n=-1.4$ power law normalized to match the present non-linear scale, the dispersion would be $\\sim 100~\\kms$. This $n$ corresponds roughly to the slope on the non-linear scale in the real $\\Lambda$CDM Universe, but $\\Lambda$CDM contains much less initial small-scale power -- not enough to bootstrap the small starting dispersion up to a significant value within linear theory (viewed very broadly, structure formation has actually taken place rather suddenly and recently, in spite of the usual ``hierarchical'' description). The next order PT calculation, which I carry out only at an order of magnitude level, should drive the dispersion up into balance with the growing structure, accounting for small dispersion effects seen recently in simulations. ", "introduction": "Observing the large-scale density fluctuations in the Universe is one of the best ways we have to approach many fundamental questions about the Universe, e.g., understanding inflation, dark matter, dark energy, the curvature of the Universe, neutrino masses, possible extra dimensions, modifications of gravity, etc. \\citep[e.g.,][]{2009ApJS..180..330K, 2005PhRvD..71j3515S, 2006JCAP...10..014S,2006PhRvL..97s1303S,2007PhRvD..76h3004S, 2003ApJ...594L..71A,2009arXiv0907.2257D,2009arXiv0908.2285L, 2009MNRAS.398..321G,2009arXiv0909.0751B,2008PhRvD..78l3534T, 2008arXiv0810.0323M,2008PhRvD..78l3519M, 2009arXiv0906.4545S, 2009arXiv0906.4548C,2009arXiv0907.5220M}. Statistics of the current density fluctuations can be used to infer statistics of the small initial perturbations from which they grew, and in turn to understand the physics of the very early Universe. Measuring the evolution of large-scale structure (LSS) over time tells us about the present matter content of the Universe and the dynamical rules its evolution follows. Before we can learn anything about fundamental properties of the Universe, however, we must be able to compute the directly observable astrophysical quantities (e.g., the CMB \\cite{2009ApJS..180..296N}, galaxy clustering statistics \\citep{2006PhRvD..74l3507T}, \\lyaf\\ absorption \\citep{2005ApJ...635..761M,2006ApJS..163...80M,2006MNRAS.365..231V, 2007PhRvD..76f3009M,2003ApJ...585...34M,2002ApJ...580...42M}, galaxy ellipticity correlations used to measure weak lensing \\citep{2006ApJ...647..116H}, galaxy cluster/Sunyaev-Zel'dovich effect (SZ) measurements \\citep{2008RPPh...71f6902A}, and possibly future 21cm surveys \\citep{2008PhRvL.100i1303C}) given a hypothetical underlying model. As observational statistics become more and more precise, with the potential to measure more and more subtle differences between models, the requirements on the phenomenological theory needed for their interpretation become correspondingly more stringent. Currently, linear-order perturbation theory \\citep{1980lssu.book.....P, 1996ApJ...469..437S,1984ApJ...284L...9K,1987MNRAS.227....1K} provides our primary means of calculating LSS observables for cosmological parameter estimation and model testing, with only ad hoc, and recently demonstrably inadequate, attempts to correct for non-linearity once it becomes non-negligible \\citep{2008MNRAS.385..830S,2006PhRvD..74l3507T,2006JCAP...10..014S, 2008JCAP...07..017H,2009ApJS..180..330K} (a vast number of papers have been written on beyond-linear calculations, but most of these are never used in parameter estimation papers). Linear theory can robustly describe observations on very large scales or at very early times, but breaks down when the perturbations become too large on a given scale. When considering gravitational evolution only (i.e., dark matter only), numerical simulations can be used to compute the fully non-linear evolution of the density field to high accuracy (with a lot of care and computer power \\cite{2008CS&D....1a5003H,2008arXiv0812.1052H,2009arXiv0902.0429H}), however, as discussed extensively in \\cite{2009JCAP...08..020M}, numerical simulations are a less than ideal tool for interpreting future precision observations, once one considers real observables which are inevitably influenced by baryonic effects (e.g., star formation, and the accompanying complication of general gas dynamics). To summarize the argument in \\cite{2009JCAP...08..020M}: Beyond linear order perturbation theory (PT) \\citep{1980lssu.book.....P,1981MNRAS.197..931J,1983MNRAS.203..345V, 1984ApJ...279..499F,1986ApJ...311....6G,1996ApJ...473..620S, 2002PhR...367....1B,2006ApJ...651..619J} should provide the primary means of interpreting very high precision future LSS data, just like linear theory has provided the primary means in the past. The range of scales over which higher order PT will be necessary (i.e., linear theory is insufficient) and sufficient (i.e., it will be accurate enough after computing a modest number of terms) will become larger as the precision of observations increases, while the chances of robustly, completely, convincingly describing the full precision of the observations using inevitably somewhat ad hoc prescriptions for star formation and related things in simulations becomes more remote. The key difference between PT and simulations is the fact that perturbative clustering can be completely described by a finite set of well-defined parameters, no matter how complicated the small-scale physics is (at least as conjectured in \\cite{2009JCAP...08..020M}), while the need for fully non-linear calculations implies that there is generally no bound on the number or form of free parameters (more or less by definition). The idea that the importance of PT relative to simulations is increasing with time may seem backwards relative to conventional wisdom, however, my argument is that this conventional wisdom was developed for the era of not very precise observations, when corrections to linear theory were already too large to be described by PT by the time they were large enough to be statistically significant, i.e., past PT work was premature. To be clear, I am not saying that simulations will not be extremely useful, only that they are unlikely to be the leading tool for extracting fundamental cosmological information from very high precision observations (much like the situation in high energy physics, where lattice QCD simulations provide much qualitative, and recently even quite precise quantitative, insight \\cite{2008Sci...322.1224D}, but high precision constraints on models are made primarily in the regime accessible to perturbation theory \\cite{2006JPhG...33....1Y}). Following the above line of reasoning, this paper is aimed at building up our ability to do calculations based on perturbation theory. It deals exclusively with the gravity-only part of the calculation, however, this should be viewed only as an intermediate goal. PT is most necessary for practical computations of observables which can not realistically be done from first principles in a simulation -- understanding how to do computations for dark matter-only is simply a prerequisite for computing these observables. This continues a recent line of work related to renormalization/resummation methods \\cite{2006PhRvD..73f3520C, 2006PhRvD..73f3519C,2007JCAP...06..026M,2007PhRvD..76h3517I, 2008ApJ...674..617T,2007PhRvD..75d3514M,2008A&A...484...79V, 2008PhRvD..77f3530M,2008MPLA...23...25M,2008PhRvD..77b3533C, 2008JCAP...10..036P, 2008PhRvD..78j3521B,2008PhRvD..78h3503B, 2009JCAP...06..017L,2009MNRAS.397.1275W,2009PhRvD..80d3531C, 2009PASJ...61..321N}. In my opinion, the key to maximizing the usefulness of all of this work is eventually connecting it to galaxy biasing and related complications \\cite{2005PhRvD..71d3511S,2006PhRvD..74j3512M, 2008PhRvD..78h3519M,2009JCAP...08..020M}. One traditional potential limitation of LSS perturbation theory, which I focus on addressing in this paper, is that it assumes the particles at a point in space all have exactly the same velocity \\cite{2009PhRvD..80d3504P}. The equations used are those of a pressureless fluid. This is often referred to as the ``single-stream approximation'', however, I will avoid this language as it assumes a certain picture of large-scale structure formation that may or may not have anything to do with reality. I say this because, for typically observed times, stream crossing is actually ubiquitous -- in fact, there are many orders of magnitude in scale of deeply non-linear structure, to the point where the initial conditions on very small scales may be effectively forgotten. Of course, the standard language implicitly means no stream crossing when the field is in some sense smoothed on the typical scale where the perturbation theory is supposed to apply. When you look at it this way, however, it is clear that, if coldness is a good effective theory, it is not simply because the particles were initially cold, there is an additional requirement that the velocity field remains effectively cold on scales just smaller than the ones of interest. To give a qualitative preview of the paper: The exact equation for the evolution of collisionless particles is the Vlasov equation for the full distribution function in 6-dimensional phase space (particle density in position and momentum space). The standard hydrodynamic equations are derived by taking moments of the Vlasov equation with respect to momentum -- the zeroth moment gives an equation for the evolution of density, the first moment gives an equation for the evolution of bulk velocity, and higher moments are normally dropped, e.g., the second moment which would describe velocity dispersion. At first glance, one might think that standard PT could be extended by simply adding the evolution equation for the 2nd and possibly higher moments, however, that turns out to be less straightforward than it sounds. Viewed conventionally, dropping these higher moments is not really an added approximation in standard Eulerian perturbation theory. The lowest order evolution equations contain only a Hubble drag term, so any small initial velocity dispersion will rapidly become even smaller. Higher order equations contain no source terms, meaning that the higher order results are always proportional to the tiny starting value. \\cite{2001A&A...379....8V} showed that even perturbation theory using the full distribution function directly leads to the same result. I will show, however, that this argument for dropping dispersion from the perturbation theory is faulty, in that, while the lowest order terms are very small, the series is very rapidly diverging, in the sense that higher order terms are increasing in size instead of decreasing. This implies that some rearrangement of the series is needed, as in a resummation or renormalization group calculation. In case this mathematical reasoning is not sufficient motivation, recently \\cite{2009PhRvD..80d3504P} computed the velocity dispersion generated in N-body simulations, finding it to be small but not completely dynamically negligible. There has been much discussion in the past about different ways of adding velocity dispersion, or more general changes to the small-scale effective theory used for PT calculations \\cite[e.g.,][]{2004PhRvD..70f4010T,2005PhRvD..71f7302R, 2006PhRvD..73b4024S,2002MNRAS.330..907M,2005A&A...438..443B, 2007JPhA...40.6849G,2009ApJ...700..705S}, however, these papers all lacked a first principles derivation, starting from the exact equations, of the model they use. This made their usefulness for precision calculations questionable, as there were always added assumptions and/or free parameters. The point of this paper is to show how to do a straightforward computation that takes us directly from the initial homogeneous, cold, starting theory to a theory with highly developed, potentially hot, small-scale structure. The rest of the paper is as follows: In \\S\\ref{seccalc} I show how to use renormalization group-inspired ideas to reintroduce velocity dispersion from first principles. In \\S\\ref{secredshiftspace} I give a short discussion of the implications of these velocity dispersion calculations for redshift-space distortions. Finally, in \\S\\ref{secdiscuss} I discuss the results. ", "conclusions": "} The deficiencies of standard perturbation theory are a lack of control of the higher moments of the velocity distribution function, beyond the pressureless fluid approximation, problems with accuracy related to the fact that PT integrals include small scales that are generally highly non-linear, even when the scales we are interested in are weakly non-linear, and, of course, simply not working on small scales where the fluctuations are large. The deficiencies of numerical simulations are speed, and the related fact that statistics (e.g., the power spectrum) cannot be computed directly, but only as averages over realizations of the random density field, which must contain many orders of magnitude more degrees of freedom (e.g., a billion) than one really needs to describe the statistic of interest (e.g., a few dozen for the power spectrum). Additionally, and probably most importantly, the cumbersome, opaque nature of simulation results is greatly exacerbated when they are used to model galaxies or other observables instead of just dark matter, while the advantage of being more or less exact for gravity alone (at least in the straightforward limits of large box size and particle density) is lost. This paper directly addresses the PT deficiency of missing higher moments of the velocity distribution function, showing how they can be re-activated and generated at a significant level, starting from first principles. The approach here may also improve PT by providing natural regulation of small-scale structure, i.e., the Jeans smoothing that arises here has the same kind of effect as the propagator resummation of \\cite{2006PhRvD..73f3520C,2006PhRvD..73f3519C}. The philosophy of this paper is that the cold ``streams'' often discussed as ``crossing'' are mythical objects -- what one sees in the real Universe is always some evolution of effectively warm (although maybe not very warm) material. The meaning of this idea is very clear for power law power spectra, where, as we saw, non-trivial effects of velocity dispersion can be computed without any initial dispersion entering the discussion. The dispersion bootstraps itself up from an arbitrarily small start. The temperature is locked into a sort of equilibrium with the growth of structure. The situation is not as clear for $\\Lambda$CDM power spectra, not so much because the $\\Lambda$CDM is cold as because there is very little small-scale power in these models, so the dispersion computed to linear order in the power spectrum remains well below the non-linear scale (although orders of magnitude larger than it would be if there was no structure). This situation will change when calculations are done to higher order, where I showed that there is enough small-scale power generated to produce dispersion that would be far too large in absence of feedback on the structure formation itself. The initial velocity dispersion should be rendered irrelevant when the system moves into a sort of self-regulating mode, like the power law example, where the velocity dispersion and growth of structure are tightly coupled by feedback between them. Ultimately, it seems likely that the best effective small-scale model for doing perturbation theory will involve some more general balance between different pieces, i.e., density, velocity divergence, dispersion, and maybe even vorticity, etc., because we know that physically this is what the small-scale structure really is, i.e., halos which can only be described as a delicate balance of the original perturbative LSS quantities. To put this another way: hopefully, when all of the relevant elements are included, there will be some fixed-point structure for small scales with a clear physical interpretation and effectively far fewer than the original number of degrees of freedom (akin to the fixed-point power law found in \\cite{2007PhRvD..75d3514M}, but with richer structure). This paper does not exactly contain useful quantitative take-home results. The results are primarily a procedure to follow for future calculations. If there is a single equation that best represents the results, it is probably Eq. (\\ref{eqkNLsigRG}), which shows how, for a power law power spectrum, the velocity dispersion tracks the non-linear scale, with Jeans filtering erasing more and more small-scale structure as the larger scale structure grows. Eqs. (\\ref{eqTnonlinear}) and (\\ref{eqtsigevolution}) are also critical, in that they show generally how homogeneous velocity dispersion is generated out of fluctuations. The key new variable, equivalent to $\\delta$ and $\\theta$, but for perturbations in velocity dispersion, is $\\pi \\propto \\partial_i \\partial_j \\sigma^{ij}$. The next step in this line of work is to derive to the next-order equations like Eq. (\\ref{eqPddevolution}-\\ref{eqPppevolution}) (but including bispectrum equations), which are needed to have any chance of properly describing $\\Lambda$CDM. Then the results can be tested by comparison to numerical simulations. While $\\Lambda$CDM is of course the ultimate goal, tests of the theoretical concepts here might be more conveniently done with power law simulations, particularly ones with relatively blue spectra, as in \\cite{2009MNRAS.397.1275W}. Any of the renormalization approaches that have been developed recently can probably be applied to the problem of renormalizing the velocity dispersion, at least in principle, i.e., resumming a set of terms that generates the dispersion, either explicitly or through a renormalization group equation. The obstacle to doing this elegantly may be the difficulty of obtaining analytic solutions as the evolution equations become more complicated (this problem of requiring analytic solutions is, I think, the primary reason to favor the ``numerical evaluation of evolution equations for statistics'' approach advocated in this paper over other, more completely analytic, recent approaches). One might ask ``why stop with the 2nd moment of the distribution function, i.e., why can we drop $q_{ijk}$ from Eq. (\\ref{eqdispersion})?'' One hope, which will need to be verified by future calculations, is that the effect of increasingly high moments on large scales may take the form of a gradient expansion, i.e., in Fourier space a series where the effects of increasingly high moments enter multiplied by increasing powers of $k$. In this case, we would expect that, on scales where the effect of the 2nd moment are already small, the effects of higher moments will be even smaller. The result for redshift-space space distortions (Eq. \\ref{eqredshiftspace}) leads to the question: Why do we use $\\vv$, and in this paper $\\sigma_{i j}$, as the variables to be solved for perturbatively, rather than, e.g., momentum $\\left(1+\\delta\\right)\\vv$ and kinetic energy $\\left(1+\\delta\\right)\\left(v_i v_j+\\sigma_{ij}\\right)$? An answer one might have considered was that velocities are needed to compute redshift-space distortions, but here we see that the most direct quantities for that are in fact momentum and energy, with additional perturbative calculations needed when starting from $\\delta$, $\\vv$, and $\\sigma_{ij}$. A change to total kinetic energy would have the potentially substantial effect of increasing the Jeans-like smoothing of the linear power, because the zero order pressure would be larger. I do not see any clear a priori reason to favor one option over another -- they are just different ways of arranging a series of terms, which should be equivalent if one could include an infinite number of terms. Note that, while the choice of variables is optional, the renormalization of the energy-related variable is not optional -- it simply makes no sense to ignore the fact that the size of terms in the series describing one of your variables is increasing rapidly, rather than decreasing, with order. There is a lot of circumstantial evidence that using total kinetic energy could be useful. The renormalization/resummation of the propagator in \\citep{2006PhRvD..73f3520C} leads to filtering much like the Jeans filtering we find due to velocity dispersion, but with scale given by the bulk velocity power spectrum. The Lagrangian PT-based scheme of \\citep{2008PhRvD..77f3530M} leads to a similar result. Another possibility to consider would be the evolution of large-scale fields with the small-scale structure explicitly averaged out, which would naturally lead to the inclusion of small-scale bulk velocities in the dispersion term \\cite{2000PhRvD..62j3501D,2002MNRAS.334..435D,2003AN....324..560D}. One might also consider using the Schr\\\"{o}dinger equation representation of the exact Vlasov equation, proposed in \\cite{1993ApJ...416L..71W}, combined with the approach of this paper. In the end, one could view the approach in this paper as a re-activation of the program under discussion in papers like \\cite{1977ApJS...34..425D}, which set out to integrate the BBGKY equations numerically. This reconsideration is timely because of several developments over the last thirty years. We now know the appropriate class of models to focus on, especially including the appropriate the initial conditions. There has been a lot of work on both perturbation theory beyond linear order, and on N-body simulations, with the limitations of each teaching us a lot about what is needed from new methods. We also have a specific calculation to focus on: the scales where baryonic acoustic oscillations (BAO) are observable \\citep{1998ApJ...496..605E, 2003ApJ...594..665B,2003PhRvD..68h3504L,2003ApJ...598..720S, 2004ApJ...615..573M,2005ApJ...631....1G, 2005MNRAS.357..429A,2005MNRAS.363.1329B,2006MNRAS.365..255B, 2007PhRvD..76f3009M}, which points us toward the perturbative approach that motivates truncating the hierarchy and believing that high precision can be achieved (in contrast to \\cite{1977ApJS...34..425D}, who were focused on the more strongly non-linear regime, where the truncation used here is not well-motivated). The same scales are also potentially the most powerful for other measurements based on, e.g., redshift-space distortions \\cite{2008arXiv0810.0323M}. The basic form of calculation I do here is completely standard in some other areas of cosmology. For example, the evolution of the homogeneous (background) value of an interacting scalar field in the early Universe is affected by quantum and thermal fluctuations. Its evolution is not described by the original equation of motion with all perturbations ignored, but by a renormalized effective potential, which is at least formally infinitely different from the original potential, i.e., completely dominated by the part due to fluctuations \\cite{2005pfc..book.....M}. Another interestingly similar calculation is the development of equilibrium after preheating after inflation. \\cite{2000hep.ph...11159F, 2001PhRvD..63j3503F, 2006JCAP...07..006D,2006PhRvD..73b3501P,2007PhRvD..75d3518F} perform fully non-linear simulations of the interaction of scalar fields, much like large-scale structure simulations, with the added twist that the background density and pressure are affected by the fluctuations. The evolution of the scale factor is calculated by taking spatial averages over the fluctuations as the simulation is running. In the beginning, the nearly homogeneous inflaton field dominates, and a very naive calculation might compute the expansion of the simulation in advance assuming homogeneous evolution, but by the end of the simulation the homogeneous component has actually practically disappeared, with the background evolution completely dominated by the fluctuations, which behave like radiation. Of course no one would ever do the very naive calculation just mentioned, where the affect of the fluctuations on the background equation of state is ignored... except that this is what we do when we assume that the tiny initial temperature of CDM means that it will forever remain cold. Substantial work remains to determine if the approach in this paper enhances the accuracy of predictions of quasi-linear clustering in the real Universe; however, it is now possible to consider an effect that was previously outside of any first-principles computational control in this kind of perturbation theory. The bottom line of this paper is that even linear theory fluctuations necessarily imply a significant one-loop renormalization of the background velocity dispersion." }, "0910/0910.3836_arXiv.txt": { "abstract": "We explore potential strategies for testing General Relativity via the coherent motions of galaxies. Our position at $z=0$ provides the reference point for distance measures in cosmology. By contrast, the Cosmic Microwave Background at $z \\simeq 1100$ acts as the point of reference for the growth of large scale structure. As a result, we find there is a lack of synergy between growth and distance measures. We show that when measuring the gravitational growth index $\\gamma$ using redshift-space distortions, typically $80\\%$ of the signal corresponds to the local growth rate at the galaxy bin location, while the remaining fraction is determined by its behaviour at higher redshifts. In order to clarify whether modified gravity may be responsible for the dark energy phenomenon, the aim is to search for a modification to the growth of structure. One might expect the magnitude of this deviation to be commensurate with the apparent dark energy density $\\Omega_\\Lambda(z)$. This provides an incentive to study redshift-space distortions (RSD) at as \\emph{low} a redshift as is practical. Specifically, we find the region around $z = 0.5$ offers the optimal balance of available volume and signal strength. ", "introduction": "Dark energy is a low-redshift phenomenon, and in accordance with the standard $\\Lambda CDM$ model, it appears to exert a rapidly decaying influence towards higher redshifts, at a rate approaching $(1+z)^{-4}$. Physical models of dark energy may be distinguished by an equation of state $w \\neq -1$, while a break from general relativity would likely exhibit a distinctive structure formation history. The consequences of a modification to general relativity are somewhat speculative at this stage, but naturally one might expect any alteration of the growth rate to become particularly prominent at late times, in accordance with the observed change in global dynamics. For instance, the $f(R)$ models explored by Hu \\& Sawicki \\cite{2007PhRvD..76j4043H} found this to be the case on large scales, as did He et al \\cite{2009JCAP...07..030H} with their interacting model. A number of probes are capable of measuring $w$ via its influence on the redshift-distance relation. It is rather more difficult to study the growth rate, due to our uncertainty in the behaviour of galaxy bias, but there are currently two promising avenues available for future exploration. Weak gravitational lensing provides a direct measurement of the dark matter distribution, and its evolution with redshift. However the focus of this work will be redshift-space distortions, which exploit the relationship between the large-scale coherent velocities of galaxies and the growth rate of perturbations. Weight functions have previously been applied to the equation of state as a function of redshift $w(z)$ \\cite{2003MNRAS.343..533D,2005PhRvD..71h3501S,2006PhRvD..73h3001S}, and this work extends the concept to the growth index $\\gamma$ \\cite{2005PhRvD..72d3529L}. They are designed to illustrate the redshift sensitivity of a given survey, and this is often quite different from the source redshift distribution. Weight functions are closely related to, and may be derived from, the principal component approach. In related work, Zhao et al. \\cite{2009arXiv0905.1326Z} recently explored the principal components of the metric ratio, a different quantity also often denoted by $\\gamma$. In \\S\\ref{sec:wts} we briefly review the concept of principal components and their application in deriving the weight function. \\S\\ref{sec:rsd} explores the RSD redshift sensitivity to the growth index $\\gamma$, and compares these weight functions to those of the dark energy equation of state. \\S\\ref{sec:opt} addresses the question of which redshift is most efficient at measuring $\\gamma$, while in \\S\\ref{sec:gro} we consider a change of parameterisation to the growth rate $f$. ", "conclusions": "At present there are two key approaches to study the evolution of dark matter perturbations, namely redshift-space distortions and weak gravitational lensing. They will provide a crucial piece of evidence in determining whether the phenomenon of dark energy may be attributed to a physical entity, or is simply due to a misunderstanding of the laws of gravity. In this work we have quantified the epoch at which a galaxy redshift survey would be sensitive to the growth index $\\gamma$. Specifically, given the absence of any weight at redshifts lower than that of the survey, low redshift surveys are left with the significant advantage of probing a broader behaviour $\\gamma(z)$. Another interesting feature we have highlighted is that approximately $80 \\%$ of the ``weight\" for $\\gamma$ for any given redshift bin corresponds to the local value of $\\gamma(z)$ at the location of the bin. The main limitation of a low redshift survey would be the available comoving volume, along with the stronger prevalence of nonlinear perturbations which may hinder an accurate determination of $\\beta$. However, for a given error on $\\beta$, one reaches a much better determination of $\\gamma$ compared to higher redshifts. By contrast, a weak lensing survey with source galaxies at any redshift will still be sensitive to the value of $\\gamma(z)$ across \\emph{all} redshifts $(02$, which we have neglected, will further advance the determination of cosmological parameters. The prospects for testing standard cosmology and in particular general relativity are promising. Improved understanding of the translinear density regime and velocities would further extend the number of usable power spectrum modes, while complementarity with other Stage IV experiments utilizing supernova distances, CMB measurements, and weak lensing data would give powerful leverage on both the gravitational growth index and other cosmological parameters. The capability of probing beyond-Einstein gravity opens up a new window for our understanding of cosmic acceleration and fundamental physics." }, "0910/0910.2417_arXiv.txt": { "abstract": "The problem of detecting dark matter filaments in the cosmic web is considered. Weak lensing is an ideal probe of dark matter, and therefore forms the basis of particularly promising detection methods. We consider and develop a number of weak lensing techniques that could be used to detect filaments in individual or stacked cluster fields, and apply them to synthetic lensing data sets in the fields of clusters from the Millennium Simulation. These techniques are multipole moments of the shear and convergence, mass reconstruction, and parameterized fits to filament mass profiles using a Markov Chain Monte Carlo approach. In particular, two new filament detection techniques are explored (multipole shear filters and Markov Chain Monte Carlo mass profile fits), and we outline the quality of data required to be able to identify and quantify filament profiles. We also consider the effects of large scale structure on filament detection. We conclude that using these techniques, there will be realistic prospects of detecting filaments in data from future space-based missions. The methods presented in this paper will be of great use in the identification of dark matter filaments in future surveys. ", "introduction": "In a universe dominated by Cold Dark Matter (CDM) such as our concordance cosmology, $\\Lambda$CDM, hierarchical structure formation takes place with the growth of collapsed objects progressing via merging of smaller objects. Simulations of structure formation show that galaxy clusters are located at the intersections of filaments, which form a web-like structure throughout the universe. The processes of merging and continuous accretion of surrounding matter are thought to occur highly anisotropically, with the filaments channeling matter along preferred directions. Modern galaxy surveys such as 2dFGRS (e.g. \\cite{Coll01}) and SDSS (e.g. \\cite{yorksdss}, \\cite{Adel08}) have provided a dramatic picture of this so-called `cosmic web', where voids lie between filamentary arms traced by galaxies. From a theoretical perspective, \\citet{Bond96} developed a theory for the cosmic web in a CDM cosmogony, where filaments are a product of a primordial tidal field evolving under non-linear effects. Detection of inter-cluster filaments has important implications for cosmology, confirming the prediction that structure grows highly anisotropically from small initial perturbations in the cosmic density field. Their detection provides an important validation of our picture of hierarchical structure formation and cluster evolution \\citep{Colfil}. Dark matter filaments also have a significant role to play when considering the total mass budget of the universe - the galaxies and gas channeled along these filaments may account for as much as half of the baryonic mass budget in the universe. Inter-cluster filaments have been the target of a number of observational searches. The earliest of these used bremsstrahlung X-ray emission to detect hot gas being channeled along filaments. For example, \\cite{Briel95} used X-ray data from the ROSAT All-Sky Survey, but failed to find evidence for filaments. Subsequent X-ray searches have also proved inconclusive (\\cite{scharf00}, \\cite{KB99}, \\cite{durret03}), the main issue being that it is difficult to ascertain whether strong X-ray emission is due to matter in a filament or due to the clusters interacting. Another proposed observational method looks for overdensities of galaxies relative to the background. \\cite{PD04} reported finding a short filament between galaxy clusters A1079/1084 using such a method. \\cite{Ebeling} report the detection of a filament leading into MACS J0717.5+3745. However, this approach is susceptible to large errors and requires spectroscopic data to establish the redshifts of the target galaxies. An alternative detection technique uses a `skeleton', which is defined as the subset of critical lines joining the saddle points of a field to its maxima while following the gradient's direction. The skeleton formalism has been used in numerical simulations to extract and analyze the filamentary structure of the universe (\\cite{Nov}, \\cite{sousbie}). Work by \\cite{Hahn} and \\cite{Aragon} demonstrates that it is possible to use the Hessian of either the potential or density field to provide a dynamical classification of filamentary structure. This work was extended by \\cite{romero}. The number of eigenvalues with magnitude above a certain threshold at each grid point determines whether that point belongs to a void, sheet or filament. \\cite{Hahn} investigated the effect filaments have on the properties of dark matter halos, and found that filaments influence the magnitude and direction of halo spin as well as halo shape. There are a number of other filament detection methods, not elaborated here, which are discussed in \\cite{Pim05}. One of the most promising methods of filament detection is weak gravitational lensing. Since gravitational lensing is insensitive to whether the matter is luminous or dark, and to its dynamical state, it is an ideal probe of dark matter. A number of attempts have been made to image filaments using weak lensing. \\citet{Gray02} claimed to detect a filament in the A901/902 supercluster, but this was later shown by \\cite{Heymans08} to be an artifact of the mass reconstruction. \\cite{Kaiser98} reported a detection of a filament between two of the galaxy clusters in the supercluster MS 0302+17, but these results could not be reproduced by \\cite{gav04}, and hence this detection is called into question. \\cite{Dietrich} used a variety of methods to detect a filament between A222/223 - however, they point out that it is difficult to objectively define whether this lensing signal is caused by a real filament. The Hubble Space Telescope Cosmic Evolution survey (COSMOS) has provided corroboratory evidence for a cosmic network of filaments \\citep{Massey}. Although there have been a number of observational attempts to detect the filamentary cosmic web with various degrees of success, conclusive evidence has proved elusive in all but a few cases, and widely applicable methods are still lacking. In this paper we assess existing techniques to detect filaments in weak lensing data and present two new, more effective, detection methods. In particular we present a method based on multipole filtering of the shear field, and a method based on Markov Chain Monte Carlo techniques that allows both detection and quantification of filament profiles. Motivated by the promise of weak lensing data from many thousands of galaxy clusters from large surveys such as the Dark Energy Survey (DES) and Large Synoptic Survey Telescope (LSST) we also consider detection in stacked data sets from a number of clusters. These methods are then applied to single and stacked synthetic weak lensing data sets from the fields of clusters from the Millennium Simulation \\citep{springel05}. We will show that weak lensing is an ideal and versatile tool to detect dark matter filaments. This paper is organized as follows. After a review of the basic lensing formalism (section 2), in section 3 we discuss the various methods by which we may detect filaments using weak lensing. These methods are then applied to data sets from the fields of simulated clusters. In section 4 we describe these simulations and in section 5 we detail our results. In section 6 we analyze the effects of large scale structure noise on our results. Finally, in section 7 we present our discussions and conclusions. ", "conclusions": "The typical strength of an intercluster filament is $\\kappa \\sim 0.01$, compared to a convergence in the central cluster regions of $\\kappa \\sim 0.5$. Therefore, the reason filaments are so difficult to detect using any method is due to the fact that they are much weaker than the cluster itself, and are quickly `lost' in the noise originating from the intrinsic ellipticity dispersion of the background galaxies. In this paper, we have shown several methods that can be used to reliably detect filaments. By implementing a variety of methods, we have also shown how versatile weak gravitational lensing is as a tool for detecting filaments - each method described in the paper has different strengths and weaknesses that make them suited to different data sets. The mass reconstruction is one of the most common methods to detect overdensities in a mass distribution. We applied the finite field mass reconstruction algorithm of \\cite{seitzrec} to both single and stacked clusters. For single clusters, reconstructions using a background galaxy density of 30 arcmin$^{-2}$ weakly detect only the strongest filaments, with the majority of filaments undetected. As background galaxy density increases, so does the ease of filament detection - at 100 arcmin$^{-2}$ we are able to identify some filaments in reconstructions. The mass reconstruction is most effective when used with stacked data sets. The stacking method has an impact on the quality of the results, with the double-cluster method proving more effective than aligning cluster major axes. A major disadvantage of mass reconstructions is that they can give false filamentary structures due to the smearing effect of the smoothing scale. This is not a problem unique to mass reconstructions, and any method which smoothes results over certain scales in an aperture will suffer from the same difficulty. This is a particular problem for closely-spaced clusters, which will appear to merge in the reconstructed image if the smoothing scale is greater than the inter-cluster separation. The second major disadvantage of a mass reconstruction is that it does not allow easy calculation of the significance of any filament detection. The origin of this effect lies in the fact that in any reconstruction, the shear field must be smoothed to avoid infinite noise \\citep{ks93}. This leads to correlated errors in the resulting convergence map, making error bars very difficult to construct. Thirdly, because filaments are weak, they are often located in regions that are dominated by noise. As \\cite{Dietrich} points out, in such a scenario the value of the mass sheet degeneracy can fluctuate by as much as the filament strength between different galaxy populations, making any filament detection in an individual cluster more questionable. This third point does not apply for the stacked clusters, as the stacking ensures that noise is reduced to sub-dominant levels even in filamentary regions. Therefore, it is advisable to use the mass reconstruction in conjunction with other methods described in this paper, both to confirm any detection and to allow the significance of any detection to be accurately assessed. In general, the multipole moments of the convergence peformed poorly, and were unable to register any strong filament detections in single clusters. Although \\cite{Dietrich} used $Q^{(2)}$ to provide evidence for a filament in A222/223, this cluster pair was very closely spaced and the corresponding filament was thus much stronger than any considered in this paper. It is also unclear how much of a contribution the filament makes to the quadrupole signal, as it is likely that the majority of the signal comes from the clusters themselves. The quadrupole signal fails to detect filaments in the stacked data due to misalignments. Misalignment between filaments will create a `mass-sheet' in the resulting stacked data, which will give zero quadrupole signal. It is not only possible to define multipole moments of the convergence, but also of the shear. Here we have shown that the shear monopole is very effective in picking out filaments. The reason for this is that the shear monopole performs a (weighted) sum of all shears in an aperture. If these shears are coherent over the aperture, then the net result will be a large signal. The shear signals corresponding to a filament will all point in approximately the same direction along the filament's length, and thus filaments are well suited to detection using this method. However, this method is not tailored to exclusively find filaments, and a variety of mass distributions will give a strong signal. The LSS adds noise to the shear monopole signal which can, in the cases of weak filaments or low galaxy densities, cause confusion, false-identification and obscuration of the filament signal. However, we have shown that at galaxy densities characteristic of future space-based missions, or if particularly strong filaments are targeted, the filament signal will still stand out clearly against the LSS noise. We demonstrated that one can use MCMC techniques to fit filament profiles, even at relatively low galaxy densities. This is an important result as it shows that weak lensing can be used not only for filament detection, but also for quantitative analysis of filament properties. In the future, this will allow us to directly compare the predictions of simulations to observations. The scatter in the best fit parameters caused by LSS has been quantified, and hinders the ability of this method to quantify profiles accurately - in future work, the LSS should be incorporated into the likelihood function to minimize the impact on the results (e.g. as per the prescription laid out in \\citet{HuWh}). The MCMC method was also shown to be quite versatile. It can be used on its own, as it was when we looked for filaments between cluster pairs, or it can be used in conjunction with other techniques such as the mass reconstruction or in particular the shear filter to first of all search for a filament candidate. There are a number of caveats associated with using MCMC fits. Firstly, it is model-dependent - this involves assuming a profile for the filament, in this paper based on the findings of \\cite{Colfil}. Models that more accurately reflect the `true' filament profiles will achieve better likelihood values. Secondly, since it is very difficult to write down a model for irregular or warped filaments, this method is only suitable for investigating the properties of straight, regular filaments. This represents only a fraction of the total filament population. Thirdly, the method works best if we have some prior information on the location of the filament, either from one of the other detection methods, or by using cluster pairs between which filaments are thought to be constrained. Without a prior on the filament position, fitting any profile would become a complicated procedure. Stacking is a technique that can be used in conjunction with any of the methods described above, to improve the chance of filament detection. Stacking is particularly useful when each individual cluster lenses a low background galaxy density, but the survey contains many such clusters. In this paper we suggested two possible stacking methods: the first aligned cluster major axes in the stacked images and the second aligned axes joining cluster pairs. The latter method is far more effective. This is because, although filaments are often observed to align with the major axis of a cluster, there is no guarantee that they are either regular or strong. In almost all cases there is imperfect alignment of the filaments with the cluster major axis, and sometimes there is no filament aligned in this direction at all. If irregular or misaligned filaments are stacked the net effect will be to produce a low density mass sheet. This is not ideal for any of the methods described in this paper. On the other hand, filaments between cluster pairs tend to be straight, strong and regular and hence are easily stacked. The only disadvantage of using cluster pairs is their relative scarcity - whereas any cluster can be used when aligning the major axes, fewer cluster pairs will exist in a survey. In summary, in this paper we have presented a variety of new detection methods, some of which (for example the shear filter and MCMC) give promising results. However, due to the low surface densities of filaments, and the additional effects of LSS noise, the task of detecting filaments in actual data remains difficult. In reality, filament detection will either require the high galaxy densities of future space-based missions, or the stacking of a number of clusters. Current space-based lensing studies achieve $\\approx$ 80 galaxies arcmin$^{-2}$ (e.g. \\citep{Leonard}), and we could achieve in excess of 250 arcmin$^{-2}$ for deep space-based observations (e.g. \\citep{Rhodes}). Although there is a trade-off between the area and depth of a survey, so that those targeting many thousands of square degrees will typically not go so deep, the most massive clusters such as those considered in this work will be prime targets for deep targeted observations. Stacking the data from the fields of less massive clusters in non-targeted large area surveys will also be extremely useful in statistically constraining the properties of filaments. There is one further method, not discussed in this paper, that could prove valuable in detecting filaments - gravitational flexion. In traditional weak lensing studies, such as the formalism used in this paper, the lens equation is approximated as linear. This means that any distortion in the ellipticity of a galaxy will be aligned purely tangentially to the lens. Physically, such a linear approximation is equivalent to assuming no variation in the lens field over the scale of the lensed image. However, if we do not make such an assumption, the lens equation becomes non-linear, and in addition to the convergence and shear, gravitational distortion is determined by the first and second flexion. Taken together, flexion introduces `arciness' to the image and some radial alignment with respect to the lens (see \\cite{GB05} for further details). It has been demonstrated, for example by \\cite{LKW09}, that gravitational flexion is an effective probe of substructure in galaxy clusters. This suggests that flexion may also be a useful tool for filament detection. Besides detecting filaments using weak lensing data, as we have noted in the Introduction, there are also a variety of other complementary detection methods. Sunyayev-Zeldovich observations could in future be used in combination with more established methods of filament detection (e.g. weak lensing, X-ray, galaxy overdensity) to provide more optimal algorithms for detection and even stronger constraints on profiles. The area of overlap between DES and the South Pole Telescope, SPT, will result in several tens of thousands of galaxy clusters in common. The ability to probe the total matter content of clusters and their environments, coupled with deep optical and infrared imaging and spectroscopic observations, will enable us to determine galaxy bias as a function of environment. It will also improve our understanding of the formation of the most massive bound objects in the universe. Weak lensing of background galaxy populations becomes less effective for use as a probe of higher redshift structures \\citep{lewking} and beyond $z\\sim 1$ an alternative source population will be essential -- such as the 21cm emission from high redshift proto-structures that will be imaged with SKA, or the CMB as seen by a future mission. The ability to trace the evolution of filaments in the cosmic web will have important consequences for models of structure formation e.g. how much material is bound in filaments as a function of epoch." }, "0910/0910.2883_arXiv.txt": { "abstract": "{Prominences are partially ionized, magnetized plasmas embedded in the solar corona. Damped oscillations and propagating waves are commonly observed. These oscillations have been interpreted in terms of magnetohydrodynamic (MHD) waves. Ion-neutral collisions and non-adiabatic effects (radiation losses and thermal conduction) have been proposed as damping mechanisms.} {We study the effect of the presence of helium on the time damping of non-adiabatic MHD waves in a plasma composed by electrons, protons, neutral hydrogen, neutral helium (\\ion{He}{i}), and singly ionized helium (\\ion{He}{ii}) in the single-fluid approximation.} {The dispersion relation of linear non-adiabatic MHD waves in a homogeneous, unbounded, and partially ion\u00a1zed prominence medium is derived. The period and the damping time of Alfv\\'en, slow, fast, and thermal waves are computed. A parametric study of the ratio of the damping time to the period with respect to the helium abundance is performed.} {The efficiency of ion-neutral collisions as well as thermal conduction is increased by the presence of helium. However, if realistic abundances of helium in prominences ($\\sim 10\\%$) are considered, this effect has a minor influence on the wave damping.} {The presence of helium can be safely neglected in studies of MHD waves in partially ionized prominence plasmas.} ", "introduction": "Small-amplitude oscillations and propagating waves are commonly observed in both quiescent and active region prominences/filaments. They have been interpreted in terms of magnetohydrodynamic (MHD) eigenmodes of the magnetic structure and/or propagating MHD waves. The reader is referred to some recent reviews for more information about the observational and theoretical backgrounds \\citep{oliverballester02,engvold04,ballester,banerjee,engvold08} Prominence oscillations are known to be quickly damped, with damping times corresponding to a few oscillatory periods \\citep[this topic has been reviewed by][]{oliver, mackay}. Several damping mechanisms of MHD waves have been proposed, non-adiabatic effects and ion-neutral collisions being the more extensively investigated. In order to understand in detail these effects, they have been studied in simple configurations such as unbounded and homogeneous media. \\citet{carbonell04} investigated the time damping in a homogeneous prominence medium taking non-adiabatic effects (optically thin radiation losses and thermal conduction) into account. Later on, the spatial damping was studied by \\citet{carbonell06} and the effect of a background mass flow was analyzed by \\citet{carbonell09}. Subsequently, some works have extended these previous results by considering the presence of the coronal medium \\citep{soler07,soler08,soler09NA}. The common conclusion of these investigations is that only slow and thermal waves are efficiently damped by non-adiabatic effects, while fast waves are very slightly damped and Alfv\\'en waves are completely unaffected. On the other hand, the influence of partial ionization on the propagation and time damping of MHD waves has been also investigated in an unbounded medium. \\citet{forteza07} followed the treatment by \\citet{brag} and derived the full set of MHD equations along with the dispersion relation of linear waves in a partially ionized, single-fluid plasma \\citep[see also][]{pinto}. The presence of electrons, protons, and neutral hydrogen atoms was taken into account, whereas helium and other species were not considered. In a subsequent work \\citep{forteza08}, they extended their previous analysis by considering radiative losses and thermal conduction by electrons and neutrals. Their main results with respect to the fully ionized case \\citep{carbonell04} were, first, that ion-neutral collisions (by means of the so-called Cowling's diffusion) can damp both Alfv\\'en and fast waves but non-adiabatic effects remain only important for the damping of slow and thermal waves, and second, that there exist critical values of the wavenumber in which the real part of the frequency vanishes, so wave propagation is not possible for larger wavenumbers. Again, applications to a more complex cylindrical geometry have been also performed \\citep{soler09IN,soler09INRA} On the basis of these previous results, it seems clear that partial ionization plays a relevant role on wave propagation in prominences. Prominences are roughly composed by 90\\% hydrogen and 10\\% helium but, to date, all the investigations considered a pure hydrogen plasma. Therefore, the effect of the presence of helium on the propagation and damping of MHD waves is still unknown and is the motivation for the present work. Here, we consider an unbounded and homogeneous prominence medium permeated by a homogeneous magnetic field. The plasma is assumed to be partially ionized, electrons, protons, neutral hydrogen, neutral helium (\\ion{He}{i}), and singly ionized helium (\\ion{He}{ii}) being the species taken into account. Recent studies by \\citet{labrosse} indicate that for central prominence temperatures, the ratio of the number densities of \\ion{He}{ii} to \\ion{He}{i} is around 10\\%, whereas the presence of \\ion{He}{iii} is negligible. This result allows us to neglect \\ion{He}{iii} in this work. Extending the works by \\citet{forteza07,forteza08}, the derivation of the basic MHD equations for a non-adiabatic, partially ionized, single-fluid plasma has been generalized by considering now five different species, allowing us to study how the presence of neutral and singly ionized helium affects their previous results. This paper is organized as follows. The description of the equilibrium and the basic equations are given in Sect.~\\ref{sec:equations}. The results are discussed in Sect.~\\ref{subsec:results}. Finally, Sect.~\\ref{sec:conclusions} contains the conclusion of this work. ", "conclusions": "\\label{sec:conclusions} In this work, we have studied the effect of helium (\\ion{He}{i} and \\ion{He}{ii}) on the time damping of thermal and MHD waves in a partially ionized prominence plasma. This is an extension of previous investigations by \\citet{forteza07,forteza08} in which helium was not taken into account. We conclude that, although the presence of neutral helium increases the efficiency of both ion-neutral collisions and thermal conduction, its effect is not important for realistic helium abundances in prominences. In addition, due to the very small \\ion{He}{ii} abundance for central prominence temperatures, its presence is irrelevant to the wave behavior. This conclusion applies both to the free propagation case and the constrained propagation by a waveguide case. Although the role of \\ion{He}{ii} (or even \\ion{He}{iii}) could be larger for typical prominence-corona transition region temperatures, the present result allows future studies of MHD waves and oscillations in prominences to neglect the presence of helium." }, "0910/0910.2551_arXiv.txt": { "abstract": "Empirical evidence for both stellar mass black holes ($\\mbh <10^{2}\\,\\msun$) and supermassive black holes (SMBHs, $\\mbh>10^{5}\\,\\msun$) is well established. Moreover, every galaxy with a bulge appears to host a SMBH, whose mass is correlated with the bulge mass, and even more strongly with the central stellar velocity dispersion $\\sigma_c$, the $\\msigma$ relation. On the other hand, evidence for \"intermediate-mass'' black holes (IMBHs, with masses in the range 100--$10^{5}\\,\\msun$) is relatively sparse, with only a few mass measurements reported in globular clusters (GCs), dwarf galaxies and low-mass AGNs. We explore the question of whether globular clusters extend the $\\msigma$ relationship for galaxies to lower black hole masses and find that available data for globular clusters are consistent with the extrapolation of this relationship. We use this extrapolated $\\msigma$ relationship to predict the putative black hole masses of those globular clusters where existence of central IMBH was proposed. We discuss how globular clusters can be used as a constraint on theories making specific predictions for the low-mass end of the $\\msigma$ relation. ", "introduction": "The empirical evidence for the ubiquity of both stellar mass black holes (black hole mass $\\mbh$ of ~\\abt~$1-15\\,\\msun$) and supermassive black holes (SMBHs, $\\mbh$~$>10^{5}~\\msun$) is well established. While it is estimated that there are about $10^7-10^9$ stellar-mass black holes in every galaxy \\citep[e.g.,][]{ShapiroTeukolsky83,BrownBethe94}, every galactic bulge appears to host a SMBH. Moreover, the kinematically determined mass of these SMBHs is correlated with the mass of the bulge, and even more strongly with the central stellar velocity dispersion $\\sigma$, the $\\msigma$ relation \\citep{FerrareseMerritt00,gebhardtetal2000c,KormendyGebhardt01,Tremaine02}. Active galactic nuclei (AGN), which have long been believed to be driven by accretion around SMBHs and whose SMBH masses have been estimated from reverberation-mapping, are also consistent with the $\\msigma$ relationship \\cite[e.g.,][and references therein]{Wandel02,Bentz09}. The corollary is that the nuclear activity is a phase (or more) in the life of (at least) every galaxy with a bulge. The $\\msigma$ relation has been extended down to black hole masses of \\abt~$10^5\\,\\msun$ \\citep{Barth05,GreeneHo06} by searching for central BHs within very low-luminosity AGN. Their dynamical studies are an unambiguous verdict on the presence of central BHs in dwarf ellipticals and very late-type spirals (e.g., NGC4395 by \\citet{GreeneHo07} and references therein). Even lower mass black holes have been inferred from non-dynamical methods for other low-luminosity AGN \\cite[e.g.,][]{Dong07}. In spite of considerable efforts, however, evidence for black holes of still lower mass, \\viz the intermediate-mass black holes (IMBHs, of masses $10^2-10^4\\msun$), is relatively sparse. Attempts to discover IMBHs in globular clusters via their X-ray emission have been on for a long time, \\cite[e.g.,][]{BahcallOstriker75}, and although the recently discovered ultra-luminous X-ray sources (ULXs) have been attributed to IMBHs \\cite[e.g.,][]{ColbertMushotsky99}, this suggestion has also been contested \\cite[e.g.,][]{Berghea08}. Extrapolating the $\\msigma$ correlation down the mass scale predicts that IMBHs can be found in stellar systems that have velocity dispersions of $<30$~km/sec, clearly pointing to the globular clusters. Observational evidence for such black holes are only a handful, however ({\\it cf.} references in Table~\\ref{tab:sampleclusters}). Furthermore, theoretical results on IMBHs in globular clusters remain ambiguous, at best. Some theories predict the necessity of most (if not all) Galactic globular clusters to host central black holes \\citep[e.g.,][]{MillerHamilton02}, while some argue for the impossibility of globular clusters to form and/or retain black holes in their cores \\citep[e.g.,][]{Favata04,KawakatuUmemura05}. The importance of investigating the possibility of globular clusters hosting IMBHs cannot be overestimated. While a central IMBH would clearly impact the evolution of the globular cluster itself, more generally, IMBHs are crucial to link the formation processeses of stellar-mass BHs and SMBHs, and could have served as seeds for the growth of SMBHs. Extending the local BH mass function to the extreme low end can help in understanding whether there is a minimum galaxy mass or velocity dispersion below which BHs are unable to form or grow \\citep[e.g.,][]{Bromley04}. Globular clusters and galactic bulges are both `hot' stellar systems, and, since all large bulge systems seem to have a central black hole, to the extent that a massive, bound globular cluster can be viewed as a ``mini-bulge\", it may be that every dense stellar system (small or large) hosts a central black hole \\citep{Gebhardt02}. BHs at the low end of the mass ladder can be used as a constraint on theoretical models with different predictions on the $\\mbh-\\s$ behaviour. For example, the prediction that $\\msigma$ relation shall substantially steepen below $\\s=150$ km/sec \\citep{Granato04} is not supported by the low-mass AGN sample \\citep{Barth05}. The question of IMBHs also has bearing on debates in cosmology, since the cosmic mass-density of IMBHs could exceed that of SMBHs ($\\O\\approx 10^{-5.7}$), and may even account for all the baryonic dark matter in the universe, with $\\O\\approx 10^{-1.7}$ \\citep{vanderMarel03}. The few recent reports on the detection of central black holes in some globular clusters seem to suggest that globular clusters do follow the $\\msigma$ relation for SMBHs \\citep{vanderMarel02,McLaughlin06, Noyola08,Lanzoni07,Gebhardt02} and that these BHs represent the `true' IMBHs in the mass range of $10^2-10^4\\,\\msun$. In this paper, we compile published discoveries of central black holes in globular clusters and investigate the consistency of the data with the $\\msigma$ and $\\mbh-$luminosity relations. Using these globular clusters as a constraint on the slope of the extended $\\msigma$ relation, we estimate the black hole masses for a sample of globular clusters that are proposed to host IMBHs and discuss the implications. ", "conclusions": "We extend the $\\msigma$ relationship for galaxies to lower black hole masses using black holes discovered in globular clusters. The reported masses of black holes in the centres of globular clusters M15, 47~Tuc, $\\omega$~Cen, NGC~6388 and G1 are consistent with the linear extrapolation of the $\\msigma$ relationship for galaxies to the low-mass end. Using this extrapolated relationship, we have estimated the masses of putative black holes in a sample of globular clusters, and find that these clusters are consistent with the $\\mbh-$luminosity relationship as well. In the extended $\\msigma$ plot, points corresponding to different types of stellar systems occupy distinct regions, suggesting that black hole masses of \\abt~few~$\\times~10^{4}~\\msun$ represent the {\\it lowest} limit for the central black holes of galaxies. Masses of the central black holes that are below this limit correspond to the globular cluster domain (keeping in mind that G1 and $\\omega$~Cen are believed to be tidally stripped dwarf galaxies and not genuine globular clusters). The consistency of black hole masses in globular clusters with the extrapolated $\\msigma$ and $\\mbh-$luminosity relationships reinforces the idea that globular clusters harbour IMBHs in their centres. The central black holes of globular clusters place even stronger constraints than low-luminosity AGNs on theoretical models of supermassive black hole growth and evolution. If black holes in globular clusters do exist, they will rule out models that predict either steepening or flattening of the $\\msigma$ relation the low-mass region. \\appendix" }, "0910/0910.0248_arXiv.txt": { "abstract": "\\begin{list}{ } {\\rightmargin 1in} \\baselineskip = 11pt \\parindent=1pc {\\small Direct measurements of the spectra of extrasolar giant planets are the keys to determining their physical and chemical nature. The goal of theory is to provide the tools and context with which such data are understood. It is only by putting spectral observations through the sieve of theory that the promise of exoplanet research can be realized. With the new {\\em Spitzer} and HST data of transiting ``hot Jupiters,\" we have now dramatically entered the era of remote sensing. We are probing their atmospheric compositions and temperature profiles, are constraining their atmospheric dynamics, and are investigating their phase light curves. Soon, many non-transiting exoplanets with wide separations (analogs of Jupiter) will be imaged and their light curves and spectra measured. In this paper, we present the basic physics, chemistry, and spectroscopy necessary to model the current direct detections and to develop the more sophisticated theories for both close-in and wide-separation extrasolar giant planets that will be needed in the years to come as exoplanet research accelerates into its future. \\\\~\\\\~\\\\~}% \\end{list} ", "introduction": " ", "conclusions": "" }, "0910/0910.2284_arXiv.txt": { "abstract": "\\fontsize{10}{10.6}\\selectfont The Serpens SMM 1 region was observed in the 6.9 mm continuum with an angular resolution of about 0\\farcs6. Two sources were found to have steep positive spectra suggesting emission from dust. The stronger one, SMM 1a, is the driving source of the bipolar jet known previously, and the mass of the dense molecular gas traced by the millimeter continuum is about 8 $M_\\odot$. The newly found source, SMM 1b, positionally coincides with the brightest mid-IR source in this region, which implies that SMM 1b is yet another young stellar object. SMM 1b seems to be less deeply embedded than SMM 1a. SMM 1 is probably a protobinary system with a projected separation of 500 AU. ", "introduction": "The Serpens dark cloud is a nearby star-forming region (see Harvey et al. 2007 and references therein). The cloud core contains a cluster of low-mass young stellar objects, and some of them are extremely young protostars (Testi \\& Sargent 1998; Harvey et al. 2007; Winston et al. 2007). The core also harbors several Herbig-Haro objects and molecular outflows (Davis et al. 1999; Hodapp 1999). The most luminous and most deeply embedded object among them is SMM 1 (Harvey et al. 1984; Casali et al. 1993; Enoch et al. 2007). SMM 1 is a Class 0 source and associated with a bipolar jet/outflow (Rodr{\\'\\i}guez et al. 1989; Curiel et al. 1996; White et al. 1995; Larsson et al. 2000). Because of the extremely high extinction, the small scale structure of SMM 1 has been accessible only with radio interferometric observations. Images of centimeter continuum revealed a well-collimated bipolar jet showing a large proper motion away from the central protostar, indicating a dynamical age of $\\sim$60 yr (Rodr{\\'\\i}guez et al. 1989; Curiel et al. 1993). Hogerheijde et al. (1999) imaged SMM 1 in the millimeter continuum and found a single compact source surrounded by an envelope of $\\sim$9 $M_\\odot$. Hogerheijde et al. (1999) also found that the molecular outflow driven by SMM 1 has a complicated structure. One of the interesting issues in the study of star formation is the multiplicity, because the majority of stars belong to multiple-star systems (Duquennoy \\& Mayor 1991). The multiplicity of young pre-main-sequence stars is higher than that of main-sequence stars (Duch{\\^e}ne et al. 2007), which implies that multiple systems form in the early stage of star formation. In fact, bright protostars, such as IRAS 16293--2422 and NGC 1333 IRAS 4A, are usually found to be multiple systems when imaged with a high angular resolution (Looney et al. 2000). Serpens SMM 1 has been considered as an interesting example of an isolated protostar that can be understood well because it is relatively bright, massive, and simple. However, there were some indications of complexity, such as the near-IR excess (Larsson et al. 2000). Therefore, it is necessary to investigate the structure of SMM 1 with a high angular resolution and a high sensitivity. In this paper, we present the results of our observations of the Serpens SMM 1 region in the 6.9 mm continuum with the Very Large Array (VLA) of the National Radio Astronomy Observatory. We describe our radio continuum observations and archival data in Section 2. In Section 3 we report the results of the continuum imaging. In Section 4 we discuss the star-forming activities in the SMM 1 region. ", "conclusions": "" }, "0910/0910.4970_arXiv.txt": { "abstract": "We report on the measurement of the physical properties (rest frame K-band luminosity and total stellar mass) of the hosts of 89 broad line (type--1) Active Galactic Nuclei (AGN) detected in the zCOSMOS survey in the redshift range $12\\times10^{11}M_{\\sun}$) the bulk of their stellar populations have been formed earlier $z_{\\rm form}\\ga 1.6$. Our results confirm previous findings by J\\o rgensen et al. This suggests that the less massive galaxies in the distant clusters have much younger stellar populations than their more massive counterparts. One explanation is that low-mass cluster galaxies have experienced more extended star formation histories with more frequent bursts of star formation with shorter duration compared to the formation history of high-mass cluster galaxies.} ", "introduction": "Clusters of galaxies at intermediate-redshift ($11$ and our understanding of the galaxy population in these high redshift clusters is limited (e.g., Rosati et al. 2004; Demarco et al. 2007; Lidman et al. 2008; Mei et al. 2009). $X$-ray surveys (e.g., Stanford et al. 2001) have detected about a dozen clusters with $z\\ge 1$. Current surveys carried out in the $X$-rays (Romer et al.\\ 2001; Stanford et al. 2006; Finoguenov et al. 2007), optical (e.g., Gladders \\& Yee 2005), infrared (Stanford et al.\\ 2005; Lawrence et al.\\ 2007), and radio (Bran\\-chesi et al.\\ 2006) provide the basis for new but small number of discoveries. Upcoming surveys using the Sunyaev-Zel'dovich effect are expected to discover between 100 to 1000 new massive galaxy clusters at high-redshifts of $z>1$ (Carlstrom et al.\\ 2002). In this context, new spectroscopic observations of clusters of galaxies play a key role for understanding galaxy evolution in dense environments. Previous studies at $z\\sim1$ concentrated on a few of the brightest (hence more massive $\\ga 2\\times10^{11}\\,M_{\\sun}$) early-type cluster members (van Dokkum \\& Stanford 2003; Holden et al.\\ 2005), because high signal-to-noise ($S/N$) spectroscopy of distant galaxies is very expensive in telescope time ($\\ga$20 hours at 8-10m facilities). In particular, these two Fundamental Plane analyses were restricted to three and four brightest early-type cluster members, respectively. At redshifts of $z\\sim1.5$, both cluster (e.g., Cucciati et al. 2006; Cooper et al. 2007) and field (e.g., Bell et al. 2004; Faber et al. 2007; Ferreras et al. 2009; Pozzetti et al. 2009) surveys show a significant increase (by about a factor of two) in the number density of early-type galaxies from high\\-er redshift to the present day. In other words, only about 50\\% of the total baryonic (stellar) mass in early-type galaxies has been generated before $z_{\\rm form}\\sim1.5$, whereas the remainder of the bulk of the stars has been assembled at much later epochs of the universe $z_{\\rm form}\\la 2$. Through an investigation of the Fundamental Plane constraints on the evolution and formation history of early-type galaxies as well as their dark matter content are possible. The Fundamental Plane (FP) is a tight linear relation in three-dimensional log-space defined by the effective (half-light) radius $r_{{\\rm e}}$, the average surface brightness within $r_{{\\rm e}}$ ($\\langle I_{{\\rm e}}\\rangle$) and the central velocity dispersion ($\\sigma$) of early-type galaxies (Djorgovski \\& Davis 1987; Dressler et al.\\ 1987; J\\o rgensen et al.\\ 1996, hereafter JFK96). This scaling relation has proven to be a powerful tool in measuring the luminosity-weighted average ages of early-type galaxies, \\newline both at local (e.g., Bender et al.\\ 1992; JFK96; Bernardi et al.\\ 2003) and at intermediate up to redshifts of $z\\sim1.2$ (e.g., J\\o rgensen et al.\\ 1999; Wuyts et al.\\ 2004; Fritz et al.\\ 2005; di Serego Alighieri et al. 2005; Treu et al.\\ 2005; J\\o rgensen et al.\\ 2006, 2007; Fritz et al.\\ 2009). The mean age of the stellar populations in these galaxies can be directly probed, because an evolution in the zero-point offset of the FP with increasing redshift can be directly related to a change in the average mass-to-light ($M/L$) ratio of the galaxies under consideration. At $z\\sim1$ our understanding of the FP is very limited. Previous works based on small number statistics at $z\\sim1.3$ found a FP relation with a large intrinsic scatter, twice as high as in the Coma cluster (van Dokkum \\& Stanford 2003; Holden et al.\\ 2005), with the properties of these early-type cluster galaxies poorly understood. This larger intrinsic scatter is in disagreement with the FP for two galaxy clusters at $z=0.8-0.9$ (J\\o rgensen et al.\\ 2006, 2007). In this high redshift domain selection effects play an important role. Therefore, both large, well defined sample sizes as well as high quality data are needed to separate observational limitations (like sample selection or cosmic variance) from the underlying physical processes in these distant galaxies (see also discussion in Fritz et al. 2009). We present the FP for 12 early-type galaxies at $z=1.013$ that are associated with the rich, massive and $X$-ray luminous cluster of galaxies \\clfort. The galaxy cluster has been observed using a combination of high $S/N$ GMOS spectroscopy at the 8m Gemini North telescope and deep imaging of the Advanced Camera for Surveys (ACS) onboard Hubble Space Telescope (HST). Our most distant sample reaches an absolute magnitude limit of $M_B\\sim -21$ mag (rest-frame), which makes this work less susceptible to selection effects. Our study of \\clfort\\ is part of the Gemini/HST Galaxy Cluster Project (J\\o rgensen et al.\\ 2005), an extensive observational program to investigate the evolution, star formation and chemical enrichment history of galaxies in rich galaxy clusters from $z=1$ to the present-day universe. Previous results on the Gemini/HST Galaxy Cluster Pro\\-ject have been published on the stellar populations of the cluster RXJ0152.7-1357 at $z=0.83$ (J\\o rgensen et al.\\ 2005), the stellar content and FP of RXJ0142.0+2131 at $z$$=$$0.28$ (Barr et al. 2005, 2006) and the FP of RXJ0152.7-1357 and RXJ1226.9+3332 at $z=0.89$ (J\\o rgensen et al.\\ 2006, 2007). Throughout this article we adopt a $\\Lambda$CDM cosmology for a flat low-density Universe with $H_0$$=$$70\\,{\\rm km\\,s^{-1}}$ ${\\rm Mpc^{-1}}$, $\\Omega_M=0.3$, and $\\Omega_{\\rm \\Lambda}=0.7$. Unless otherwise noted, all magnitudes are given in the Vega system. ", "conclusions": "Galaxies in clusters at intermediate redshift are useful pro\\-bes to study the formation and the evolution of early-type galaxies. By measuring the kinematics and structure parameters of these galaxies their evolution in mass and mass-to-light ($M/L$) ratios can be derived. This allows us to put constraints on the formation epoch and subsequent evolution of spheroidal galaxies up to the present-day as well as to critically test the models of galaxy formation and evolution. In this paper we have presented the results of a detailed study of early-type galaxies in the distant clusters of galaxies \\clfort\\ at $z=1.013$. Based on HST/ACS imaging, surface brightness profiles, morphologies and stru\\-ctural parameters were derived for 30 galaxies within the HST/ACS field-of-view of the galaxy cluster. We have constructed the first detailed Fundamental Plane (FP) of 12 clu\\-ster early-type galaxies at redshift $z=1$. This analysis is based on a sample size that represents a factor of 3 or more improvement compared to previous studies. We have combined our distant cluster sample with our previous detailed spectroscopic study of 38 early-type galaxies in two distant galaxy clusters at $z=0.8-0.9$, RXJ0152.7-1357 and RXJ1226.9+3332 (J\\o rgensen et al.\\ 2006, 2007). This allows us to study in more detail the early-type galaxy population in high redshift galaxy clusters up to redshift unity. Compared to the local reference, the distant cluster galaxies follow a steeper FP relation and $M/L$-mass relation. This slope change can be interpreted as a difference in the epoch of the last major star formation episode for galaxies with different stellar masses. The least massive galaxies ($M=2\\times10^{10}M_{\\sun}$) in our sample have experienced their most recent major star formation burst at $z_{\\rm form}\\sim 1.1$, whereas for massive galaxies ($M>2\\times10^{11}M_{\\sun}$) the majority of their stellar populations have been formed earlier $z_{\\rm form}\\ga 1.6$. The dependence of the formation epoch on the mass confirms previous results by J\\o rgensen et al.\\ (2006, 2007). A similar mass-dependent evolution with more extended SFH has recently been found for field early-type galaxies (Fritz et al. 2009). Our results can be understood if less massive cluster galaxies have on average younger stellar populations. The distant galaxy population in these low-mass galaxies could have been built up over longer time scales and over more extended star formation histories of shorter duration but more frequent star formation bursts, compared to the formation history of the high-mass galaxy cluster counterparts. Our results support a galaxy formation scenario according to the \\textit{down-sizing} picture (Cowie et al. 1996), where the mass of galaxies hosting star formation processes decreases with the age of the Universe. A combined analysis of the FP and the absorption line indices for \\clfort\\ as well as RXJ0152.7-1357 and RXJ1226.9+3332 will be presented in J\\o rgensen et al.\\ (2009)." }, "0910/0910.4013_arXiv.txt": { "abstract": "We use a high-resolution $N$-body simulation to investigate the influence of background galaxy properties, including redshift, size, shape and clustering, on the efficiency of forming giant arcs by gravitational lensing of rich galaxy clusters. Two large sets of ray-tracing simulations are carried out for 10 massive clusters at two redshifts, i.e. $z_{\\rm l} \\sim 0.2$ and $0.3$. The virial mass ($M_{\\rm vir}$) of the simulated lens clusters at $z\\sim0.2$ ranges from $6.8\\times10^{14} h^{-1} {M_{\\odot}}$ to $1.1\\times 10^{15} h^{-1} M_{\\odot}$. The information of background galaxies brighter than 25 magnitude in the $I$-band is taken from Cosmological Evolution Survey (COSMOS) imaging data. Around $1.7\\times 10^5$ strong lensing realizations with these images as background galaxies have been performed for each set. We find that the efficiency for forming giant arcs for $z_{\\rm l}=0.2$ clusters is broadly consistent with observations. % Our study on control source samples shows that the number of giant arcs is decreased by a factor of 1.05 and 1.61 when the COSMOS redshift distribution of galaxies is adopted, compared to the cases where all the galaxies were assumed to be in a single source plane at $z_{\\rm l}=1.0$ and $z_{\\rm l}=1.5$, respectively. We find that the efficiency of producing giant arcs by rich clusters is weakly dependent on the source size and clustering. Our principal finding is that a small proportion ($\\sim 1/3$) of galaxies with elongated shapes (e.g. ellipticity $\\epsilon=1-b/a>0.5$) can boost the number of giant arcs substantially. Compared with recent studies where a uniform ellipticity distribution from 0 to 0.5 is used for the sources, the adoption of directly observed shape distribution increases the number of giant arcs by a factor of $\\sim2$. Our results indicate that it is necessary to account for source information and survey parameters (such as point-spread-function, seeing) to make correct predictions of giant arcs and further to constrain the cosmological parameters. ", "introduction": "Giant arcs are spectacular examples of strong gravitational lensing found in rich galaxy clusters. Background galaxies are stretched into long, thin arcs by the intense foreground gravitational field. Hundreds of giant arcs have been found in both optically-selected and X-ray selected clusters (e.g. \\citealt{Luppino99, ZG03, Gladders03, Sand05}). Massive clusters are efficient producer of giant arcs and the number of giant arcs is a good indicator of the abundance of massive clusters. The halo mass function is very sensitive to the cosmological parameters, especially at the massive end. Moreover, the internal structures of massive clusters, such as substructures and ellipticity, also depend on the cosmological parameters. They all affect the lensing probability (optical depth, e.g., \\citealt{Mene03a, Mene03b, Wambsganss04, Torri04, Dalal04, Li05, OLS03, Oguri08, Hilbert07, Puchwein05, Puchwein09}) in various degree. Therefore, the observation of giant arcs is a useful probe of the cosmological model, in particular the matter power-spectrum normalization, $\\sigma_8$ \\citep{Li06b}, the matter density, $\\omega0$, and to a less extent the cosmological constant, $\\Omega_{\\Lambda,0}$. However, in order to use the observations of giant arcs to constrain the cosmological models, one has to thoroughly understand how the lensing probability is affected by the distribution of background source galaxies, as well as the intrinsic properties of lens population. It has been shown that the lensing probability increases significantly with the increase of the source redshift, thus it is necessary to quantify the redshift distribution of source galaxies in lensing studies \\citep{Wambsganss04}. The lensing efficiency of massive clusters does not depend on the source size significantly \\citep{Li05}, although not all the real source information has been used, in particular their shapes. \\citet{Horesh05} first adopted realistic galaxy images with known photometric redshifts from the Hubble Deep Field (HDF) as background sources. They selected clusters from a cosmological $N$-body simulation in a $\\sim140 h^ {-1}{\\rm Mpc}$ box at $z_{\\rm l}\\sim 0.2$. The mass of the simulated lens clusters ($0.54$--$1.1\\times10^{15} h^ {-1}{\\rm M_{\\odot}}$) is similar to that of 10 X-ray--selected clusters \\citep{Smith05}, the (mean) mass range of which is around $6.3\\times10^{14} h^{-1} M_\\odot \\leq M_{200}\\leq2.0\\times10^{15} h^{-1} M_{\\odot}$. Note that the mass is calculated based on the X-ray luminosity by using the $L_{\\rm x}$-$M_{200}$ relation \\citep{Reiprich02}. They argued that the probability of producing giant arcs is $\\sim 1$ per cluster after observational effects are included, and emphasize that the number of giant arcs produced by the simulated clusters in their adopted cosmology (with $\\omega0=0.3, \\Omega_{\\Lambda,0}=0.7$ and the power-spectrum normalization of $\\sigma_8=0.9$) is consistent with the one observed in the massive clusters at redshift $0.17110^{18}$ G, the dependence of both $\\zperp$ and $\\zpara$ on the field is strong, and they exhibit de Haas-van Alphen-type oscillations. With increasing magnetic field, the amplitude of these oscillations increases, which eventually leads to negative $\\zperp$ in some regions of parameter space. We show that the change of sign of $\\zperp$ signals a hydrodynamic instability. As an application, we discuss the effects of the new bulk viscosities on the r-mode instability in rotating strange quark stars. We find that the instability region in strange quark stars is affected when the magnetic fields exceeds the value $B= 10^{17}$ G. For fields which are larger by an order of magnitude, the instability region is significantly enlarged, making magnetized strange stars more susceptible to $r$-mode instability than their unmagnetized counterparts. ", "introduction": " ", "conclusions": "" }, "0910/0910.1894_arXiv.txt": { "abstract": "We present results of optical spectroscopic and $BVR_\\mathrm{C}I_\\mathrm{C}$ photometric observations of 77 pre-main sequence (PMS) stars in the Cepheus flare region. A total of 64 of these are newly confirmed PMS stars, originally selected from various published candidate lists. We estimate effective temperatures and luminosities for the PMS stars, and comparing the results with pre-main sequence evolutionary models we estimate stellar masses of 0.2--2.4\\,M$_{\\sun}$ and stellar ages of 0.1--15~Myr. Among the PMS stars, we identify 15 visual binaries with separations of 2--10~arcsec. From archival IRAS, 2MASS, and Spitzer data, we construct their spectral energy distributions and classify 5\\% of the stars as Class~I, 10\\% as Flat SED, 60\\% as Class~II, and 3\\% as Class~III young stellar objects (YSOs). We identify 12 CTTS and 2 WTTS as members of NGC~7023, with mean age of 1.6~Myr. The 13 PMS stars associated with L\\,1228 belong to three small aggregates: RNO~129, L\\,1228\\,A, and L\\,1228\\,S. The age distribution of the 17 PMS stars associated with L\\,1251 suggests that star formation has propagated with the expansion of the Cepheus flare shell. We detect sparse aggregates of $\\sim$6--7~Myr-old PMS stars around the dark clouds L\\,1177 and L\\,1219, at a distance of $\\sim 400$\\,pc. Three T~Tauri stars appear to be associated with the Herbig Ae star SV~Cep at a distance of 600~pc. Our results confirm that the molecular complex in the Cepheus flare region contains clouds of various distances and star forming histories. ", "introduction": "\\label{Sect_1} Star forming regions in our Galactic environment offer a unique opportunity for understanding the birth of planetary systems. There is growing evidence that not only the stellar content and structure of young stellar groups, but also the evolution of protostellar envelopes and accretion disks depend on the star forming environment \\citep[e.g.][]{Hester05,Dullemond06}. Cross-comparing various star forming environments is important for understanding the origins and early evolution of stars and planets. In this paper we present our results of spectroscopic and photometric studies of a population of pre-main sequence stars associated with the molecular clouds of the Cepheus flare, located at 200--400~pc from the Sun. A detailed review on the ISM and young stellar populations in the Cepheus flare region can be found in \\citet*{Kun08}. Here we briefly review the highlights that are relevant for this study. The Cepheus flare \\citep{Hubble} molecular cloud complex consists of a large amount of dense ISM between $100\\degr \\leq l \\leq 120\\degr$ and above $b \\geq 10\\degr$ \\citep*{LyndsD,TDS, CB88,DUK}. Though the CO maps of the region presented by \\citet{Lebrun} and \\citet{GLADT} indicate a coherent cloud complex, there is evidence for multiple cloud layers over the Cepheus flare region. For instance, stars illuminating reflection nebulae or displaying 100\\,\\micron \\ excess due to their dusty environment, can be found at various distances \\citep*[e.g.][]{Racine,KVSz}, suggesting the presence of more than one dust layer between about 200 and 800~pc. Wolf diagrams of the region indicate extinction layers at 200, 300, and about 450~pc \\citep{Kun98}. Radial velocity measurements of the interstellar matter suggest the presence of several gas layers \\citep[e.g.][]{Heiles,YDMOF}. \\citet{Olano} modeled the space distribution and kinematics of the interstellar neutral and molecular gas by an expanding shell, centered on $(l,b)\\approx (120\\degr,+17\\degr)$, at a distance of about 300~pc. The presence of the {\\em Cepheus flare Shell\\/} (CFS), an old supernova remnant, readily explains the divergent cloud distances and the wide range of radial velocities. The giant radio continuum feature {\\em Loop~III} \\citep{Berk73, Spoelstra}, conspicuous in the {\\it WMAP\\/} K-band polarization map \\citep{Page07}, is nearly concentric with the molecular shell. Several giant far-infrared loops, possibly old supernova shells, have been identified in the IRAS images of the Cepheus flare region \\citep*{KMT}. Signposts of low-mass star formation in the Cepheus flare region have been detected in the molecular clouds L\\,1082 (B\\,150, GF\\,9), L\\,1148/L\\,1158, L\\,1172/L1174 (associated with the young cluster NGC\\,7023), L\\,1228, L\\,1177 (CB 230), L\\,1219 (B\\,175), L\\,1221, L\\,1251, and L\\,1261 (CB~244) \\citep[see review of ][]{Kun08}. Recently, \\citet{Kirk09} presented a list of 133 young stellar objects and candidates, based on Spitzer observations of L\\,1148/L\\,1158, L\\,1172/L1174, L\\,1228, L\\,1221, and L\\,1251. Independent distance estimations of individual clouds suggest various distances between 150 and 500~pc, but are clustered around 300\\,pc. The young open cluster NGC\\,7129, associated with the dark cloud L\\,1181, is projected on the Cepheus flare region at ($l$,$b$)=(105.4,+9.9), but is more distant than the bulk of clouds (800--1250~pc, see \\citet{Kun08}. The central region of the complex is lacking signatures of active star formation. The physical connection between the star-forming clouds remains rather uncertain. Scarce information is available on the pre-main-sequence (PMS) stellar population of the region. \\citet{HBC} list only 14 young objects in the Cepheus flare region: RNO\\,124 and PV Cep (associated with the cloud complex L\\,1148/L\\,1158), HD\\,200775, EH Cep, FU Cep, FV Cep (members of NGC\\,7023), Lk\\ha~234, BD\\,+65\\degr1637, V350~Cep, [SVS76]~NGC\\,7129~6 (members of NGC\\,7129), AS~507 (associated with L\\,1261), and the apparently isolated Herbig Ae/Be stars BH Cep, BO Cep, and SV Cep. Six further stars in NGC\\,7129 (MEG\\,1--MEG\\,3, [HL85]\\,14: \\citealt{MEG93}; GGD~33a: \\citealt{MEFG94}, and V391~Cep: \\citealt{Semkov93}) show T Tauri-like emission spectra, but no spectral types have been given for them in the literature. \\citet{Tachihara05} identified 16 weak-line T~Tauri stars in the region, by spectroscopic and photometric follow-up studies of X-ray emitting stars found in the ROSAT database. These stars are not obviously associated with any dark cloud. Understanding star formation in each dark cloud and relating star formation rates across the dark clouds requires a comprehensive deep survey. Here, we use optical and infrared observations to develop a better picture of star formation in this region. To constrain the nature of the candidate young stars, we combine new optical spectroscopy and BVRI photometry with archival infrared data from IRAS, 2MASS, and Spitzer. We describe our observations and data reduction in Sect.~\\ref{Sect_obs}. We use these data to estimate ages, luminosities, masses, and disk properties of individual young stars and to analyze the distributions of these physical parameters among each cloud. Section~\\ref{Sect_res} contains the main results, namely the HR diagrams and spectral energy distributions of the target stars. We discuss the implications of our results for the star formation history of the region in Sect.~\\ref{Sect_disc}. ", "conclusions": "From spectroscopic and photometric observations, we estimated the effective temperatures and luminosities of 78 PMS stars in the Cepheus flare region, 65 of which are newly confirmed PMS stars. Comparing the results with pre-main sequence evolutionary models we estimate stellar masses of 0.2--2.4\\,M$_{\\sun}$ and stellar ages of 0.1--15~Myr. From archival IRAS, 2MASS, and Spitzer data, we constructed the spectral energy distributions; 5\\% of the stars are Class~I, 5\\% Flat, 81\\% Class~II, and 9\\% Class~III. Star forming clouds and PMS stars of the region belong to the complexes as follows. \\begin{itemize} \\item Star formation associated with the large expanding system of Cepheus flare Shell/Loop~III occurs in the clouds L\\,1148/L\\,1158, L\\,1172/L\\,1174, L\\,1251, L\\,1228 and L\\,1261/L\\,1262. The latter two clouds are located on the approaching side of the expanding shell. \\item A sparse association consisting of 6--7~Myr old PMS stars at the low-latitude part of the region ($105\\degr \\lesssim l \\lesssim 112\\degr$, $10\\degr\\lesssim b \\lesssim 14\\degr$) can be found at a distance of 380--400\\,pc. The shapes of the clouds in this part of the region suggest shocks from the direction of the Galactic plane. \\item Low-mass PMS stars are associated with the Herbig~Ae star SV~Cep, located at $\\sim$600\\,pc. \\item NGC\\,7129, at $\\sim$800\\,pc from the Sun, is probably associated with the Cepheus Bubble, created by the older subgroup of the association Cep~OB2. Stars have been formed in at least two clumps of its natal cloud L\\,1181. \\end{itemize} The A1-type \\ha\\ emission star associated with IRAS~22219+7901 is probably more distant than the Cepheus flare clouds. Its nature remains uncertain." }, "0910/0910.3069_arXiv.txt": { "abstract": "Two recent papers (Ghez et al. 2008, Gillessen et al. 2009) have estimated the mass of and the distance to the massive black hole in the center of the Milky Way using stellar orbits. The two astrometric data sets are independent and yielded consistent results, even though the measured positions do not match when simply overplotting the two sets. In this letter we show that the two sets can be brought to excellent agreement with each other when allowing for a small offset in the definition of the reference frame of the two data sets. The required offsets in the coordinates and velocities of the origin of the reference frames are consistent with the uncertainties given in Ghez et al. (2008). The so combined data set allows for a moderate improvement of the statistical errors of mass of and distance to Sgr~A*, but the overall accuracies of these numbers are dominated by systematic errors and the long-term calibration of the reference frame. We obtain $R_0 = 8.28 \\pm 0.15 |_\\mathrm{stat} \\pm 0.29|_\\mathrm{sys}\\,$kpc and $M_\\mathrm{MBH} = 4.30 \\pm 0.20 |_\\mathrm{stat} \\pm 0.30|_\\mathrm{sys} \\times 10^6\\,M_\\odot$ as best estimates from a multi-star fit. ", "introduction": "The motions of stars in the immediate vicinity of Sgr~A* have been tracked since 1992 at the NTT/VLT and since 1995 at the Keck telescope \\citep{Eckart:1996p163,Ghez:1998p118}. With the detection of accelerations \\citep{Ghez:2000p138} and the determination of the first orbit \\citep{Schodel:2002p153} these measurements provided firm evidence for the existence of a massive black hole (MBH) at the center of the Milky Way, coincident with the radio-source Sgr~A*. The stars are used as test particles for the gravitational potential in which they move. Particularly important is the bright ($m_\\mathrm{K}\\approx14$) star S2 (S0-2 in the Keck nomenclature) orbiting the MBH in 15.9 years on an ellipse with an apparent diameter of $\\,\\approx0.2''$. Together with radial velocity data, the astrometric data allow for a geometric determination of $R_0$, the distance to the Galactic Center \\citep{Salim:1999p1362,Eisenhauer:2003p128}. The number of orbits has increased to more than 20 since then \\citep{Ghez:2005p1681,Eisenhauer:2005p117,Gillessen:2009p1117} and in particular the full S2 orbit has recently been used for improved determinations of the mass and of $R_0$ \\citep{Ghez:2008p945,Gillessen:2009p1117}. These two astrometric data sets are independent since they were obtained at different telescopes and were analyzed by different teams using different tools. On the other hand, the radial velocity data attached to the astrometric data sets largely overlap, mainly because this technically is straight-forward. The radial velocities refer to the local standard of rest and hence the inclusion of other data into a given data set needs no special care. The situation is different for astrometric data, for which only an approximate realization of an absolute reference frame currently is possible \\citep{Reid:2007p169}. This means that the exact definition of coordinates is a matter of the respective data analysis, and simply merging two lists of astrometric positions will fail (see figure~\\ref{f1}). The two groups \\citep{Ghez:2008p945,Gillessen:2009p1117} come to very similar conclusions for the mass of Sgr~A* and $R_0$. From VLT data \\cite{Gillessen:2009p1117} derived (when using the S2 orbit only) \\begin{eqnarray} R_0&=&8.31 \\pm 0.33|_\\mathrm{stat} \\,\\mathrm{kpc}\\nonumber\\\\ M_\\mathrm{MBH} &=& 4.29 \\pm 0.07|_\\mathrm{stat} \\pm 0.34|_{R_0}\\,M_\\odot\\,\\, \\end{eqnarray} where the statistical error on the mass is for a fixed distance and the second error term is due to the statistical error on $R_0$. The latter includes already the coordinate system uncertainty. In addition, a systematic error of $\\Delta R_0 \\approx \\pm^{\\,0.5}_{\\,1.0}\\,$kpc is present, owing to the uncertainties of a) how much the 2002 data (close to periastron) can be trusted and b) the assumption that the effective potential is Keplerian. The systematic error of $R_0$ also influences $M_\\mathrm{MBH}$ since the two quantities are correlated. \\cite{Ghez:2008p945} obtained from the Keck data set (neglecting 2002 data, and fixing the radial velocity $v_z$ of the MBH to 0) \\begin{eqnarray} R_0&=&8.4 \\pm 0.4\\,\\mathrm{kpc} \\nonumber\\\\ M_\\mathrm{MBH} &=& 4.5 \\pm 0.4\\,M_\\odot\\,\\,. \\end{eqnarray} The errors are of statistical nature and the mass error includes the uncertainty induced by the error on $R_0$. Probably, also a systematic error should be taken into account here. Leaving $v_z$ free yielded somewhat larger statistical errors, e.g. $\\Delta R_0=0.6\\,\\mathrm{kpc}$. The aim of this letter is to investigate the combined data set. We show that not only the fit results are in excellent agreement, but also that the combined data can be described very well by a single Keplerian orbit. The such combined data set also yields an improvement in the statistical fit errors for the derived parameters. This can be understood qualitatively by inspecting the data: Before 2002, the Keck measurements appear superior to our NTT/VLT data set. At that time, our group was using the NTT with an aperture of $3.6\\,$m, which is notably smaller than $10\\,$m for the Keck telescopes. After 2002, the VLT data has comparable errors to the Keck data for individual data points but a much denser time sampling. This is owed to the location on the southern hemisphere of the VLT and the generous allocation of telescope time to the Galactic Center projects. Hence, by suitably combining the two data sets, one should obtain a data set which combines the respective advantages.\\footnote{We use data collected between 1992 and 2009 at the European Southern Observatory, both on Paranal and La Silla, Chile; in particular from the Large Programs 179.B-0261 and 183.B-0100.} ", "conclusions": "Since more than 10 years two groups have assembled independently astrometric data sets for the stars in the central arcsecond of the Milky Way. The data sets are truly independent. They were obtained at different telescopes with different instruments. The analysis chains used different tools (deconvolution in the NTT/VLT case, PSF-fitting for the Keck data), and the definition of astrometric coordinates is implemented in different ways. While it was reassuring, that \\cite{Ghez:2008p945} and \\cite{Gillessen:2009p1117} concluded very similarly for $M_\\mathrm{MBH}$ and $R_0$, it was not clear that the agreement actually also holds for the underlying data. We were able to show that indeed the most simple assumption - namely that the two data sets only differ in the coordinate system definition - is sufficient to perfectly map the two data sets on top of each other. We presented for the first time a combined orbit fit, which however only moderately improves the accuracy by which $M_\\mathrm{MBH}$ and $R_0$ can be derived from the S2 data: \\begin{eqnarray} R_0 &=& 8.34 \\pm 0.27 |_\\mathrm{stat} \\pm 0.5|_\\mathrm{sys}\\,kpc \\nonumber\\\\ M_\\mathrm{MBH} &=& 4.40 \\pm 0.27 |_\\mathrm{stat} \\pm 0.5|_\\mathrm{sys} \\times 10^6\\,M_\\odot\\,\\,. \\end{eqnarray} For the sake of completeness, we cite also the updated numbers for a fit using not just the combined S2 data, but in addition S1, S8, S12, S13 and S14, the same stars \\cite{Gillessen:2009p1117} had used: \\begin{eqnarray} R_0 &=& 8.28 \\pm 0.15 |_\\mathrm{stat} \\pm 0.29|_\\mathrm{sys}\\,kpc \\nonumber\\\\ M_\\mathrm{MBH} &=& 4.30 \\pm 0.20 |_\\mathrm{stat} \\pm 0.30|_\\mathrm{sys} \\times 10^6\\,M_\\odot\\,\\,. \\end{eqnarray}\\\\ Besides the expected improvement in $\\Delta R_0|_\\mathrm{stat}$ from $0.17\\,$kpc to $0.15\\,$kpc, also the systematic errors have decreased mildly compared to the previous work due to the smaller influence of the S2 2002 data. We conclude that the combination does not help much in overcoming current limitations. The true value of the two data sets actually is their independence, allowing for cross checks and lending credibility to the results. Further substantial improvements in measuring the gravitational potential from Sgr~A* by means of stellar orbits probably will come from improved astrometry on existing data, from longer time lines with the existing instruments and finally from advances in instrumentation." }, "0910/0910.4964_arXiv.txt": { "abstract": "% The relation between the globular cluster luminosity function (GCLF, ${\\rm d}N/{\\rm d}\\log{L}$) and globular cluster mass function (GCMF, ${\\rm d}N/{\\rm d}\\log{M}$) is considered. Due to low-mass star depletion, dissolving GCs have mass-to-light ($M/L$) ratios that are lower than expected from their metallicities. This has been shown to lead to an $M/L$ ratio that increases with GC mass and luminosity. We model the GCLF and GCMF and show that the power law slopes inherently differ (1.0 versus 0.7, respectively) when accounting for the variability of $M/L$. The observed GCLF is found to be consistent with a Schechter-type initial cluster mass function and a mass-dependent mass-loss rate. ", "introduction": "Even though globular cluster systems (GCSs) are considered to have formed during mergers, the shape of the globular cluster luminosity function (GCLF, ${\\rm d}N/{\\rm d}\\log{L}$) differs fundamentally from that of young massive clusters (YMCs) in merging galaxies. The luminosity function of YMCs follows a power law with index $-2$ over the full mass range down to a few $100~\\msun$, while the GCLF peaks at $\\sim 10^5~\\lsun$. This has been attributed to the ongoing tidal disruption of globular clusters (GCs) and the resulting destruction of low-mass, faint GCs \\citep{fall01,vesperini01}. In studies considering this mechanism, the mass-loss rate of GCs is considered to be independent of their mass (corresponding to a disruption time $t_{\\rm dis}\\propto M$) and the mass-to-light ratio ($M/L$) is assumed to be constant \\citep[e.g.,][]{fall01,jordan07}. In order to trace back the merger history of galaxies using their globular cluster mass function (GCMF, ${\\rm d}N/{\\rm d}\\log{M}$), it is essential to obtain an accurate description for its evolution. Although the above studies reproduced the peaked shape of the GCMF, they did not account for two aspects of GC evolution: \\begin{itemize} \\item[(1)] The mass-loss rate of GCs does depend on cluster mass \\citep{baumgardt01,baumgardt03,lamers05,larsen09}, corresponding to $t_{\\rm dis}\\propto M^\\gamma$ with $\\gamma\\sim 0.7$ (see Eq.~\\ref{eq:dmdt}). This is due to the nonlinear scaling of the disruption time with the half-mass relaxation time \\citep[$t_{\\rm dis}\\propto t_{\\rm rh}^{0.75}$,][]{portegieszwart98,fukushige00}. A lower value of $\\gamma$ means that the dissolution rate of low-mass clusters is slowed down with respect to massive clusters. \\item[(2)] The $M/L$ ratio of GCs is not constant \\citep{mandushev91,baumgardt03,kruijssen08,kruijssen09} due to the preferential ejection of faint, low-mass stars from dissolving GCs. Because low-mass GCs have on average lost a larger fraction of their initial masses, $M/L$ increases with mass or luminosity. \\end{itemize} The effect of these aspects of GC evolution on the relation between the GCLF and the GCMF has been considered by \\citet{kruijssen09b}. The slope of the disruption-dominated low-mass side of the GCMF is always equal to $\\gamma$ \\citep{fall01,lamers05}. A mass-dependent mass-loss rate ($\\gamma=0.7$) therefore yields a different GCMF slope than a mass-independent mass-loss rate ($\\gamma=1$). The observed slope of the GCLF $\\sim 1$, which is seemingly consistent with the latter if a constant $M/L$ ratio is assumed (see Fig.~\\ref{fig:histcan}). However, the trend of increasing $M/L$ ratio with luminosity implies that this is not necessarily true because the slopes of the GCLF and the GCMF fundamentally differ. \\citet{kruijssen09b} have shown that the observed GCLF is in fact consistent with a GCMF with a low-mass slope of $\\sim 0.7$ when accounting for the trend of $M/L$ ratio with luminosity. \\begin{figure} \\center\\resizebox{12.4cm}{!}{\\plottwo{kruijssenj3fig1a.ps}{kruijssenj3fig1b.ps}} \\caption[]{\\label{fig:histcan} {\\it Left}: The observed GCLF of Galactic GCs \\citep{harris96}. {\\it Right}: The inferred GCMF of Galactic GCs (histogram) using $M/L_V=3$ \\citep[as in][]{fall01}. Overplotted is our model MF with a mass-dependent mass-loss rate (solid line, see Section~\\ref{sec:model}) adopting a dissolution timescale $t_0=1.6$~Myr. The dashed line shows the model for a cluster mass-independent mass-loss rate \\citep[as in][]{fall01}. Error bars are $1\\sigma$ Poissonian.} \\end{figure} We revisit the calculations of \\citet{kruijssen09b} and include a more detailed model for the evolution of the stellar mass function (SMF) within a star cluster. Previously, the low-mass star depletion was approximated by increasing the lower stellar mass limit of the SMF. In reality, the SMF evolves more gradually. This has been included in new models of the evolution of the SMF in dissolving star clusters \\citep{kruijssen09c}, which are used here. ", "conclusions": "Using a realistic model for the evolution of the SMF in GCs, we have shown that the slopes of the GCLF and the GCMF differ. As GCs dissolve, their masses decrease more rapidly than their luminosities due to the ejection of faint, low-mass stars. As a result, the slope of the GCLF is $\\sim 1$ for luminosities below the peak luminosity, while the slope of the GCMF is $\\sim 0.7$ for masses below the peak mass. This is consistent with a mass-loss rate that depends on cluster mass and on the environment. The new cluster models \\citep{kruijssen09c}, in which we account for the changing slope of the stellar mass function rather than shifting the lower stellar mass limit, show that the results from \\citet{kruijssen09b} also hold when more detailed models are applied. This substantiates that care should be taken when comparing GCLFs and GCMFs. Because of the relatively low number of GCs for which dynamical masses have been determined, we argue that the best method to compare them would be to model the GCLF rather than the GCMF, while accounting for the variability of the $M/L$ ratio. This would allow for a more accurate interpretation of the GCLF when studying correlations between its properties and the history of its galaxy." }, "0910/0910.3858_arXiv.txt": { "abstract": "{Flow-induced instabilities of magnetic flux tubes are relevant to the storage of magnetic flux in the interiors of stars with outer convection zones. The stability of magnetic fields in stellar interiors is of importance to the generation and transport of solar and stellar magnetic fields. } {We consider the effects of material flows on the dynamics of toroidal magnetic flux tubes located close to the base of the solar convection zone, initially within the overshoot region. The problem is to find the physical conditions in which magnetic flux can be stored for periods comparable to the dynamo amplification time, which is of the order of a few years.} {We carry out nonlinear numerical simulations to investigate the stability and dynamics of thin flux tubes subject to perpendicular and longitudinal flows. We compare the simulations with the results of simplified analytical approximations.} {The longitudinal flow instability induced by the aerodynamic drag force is nonlinear in the sense that the growth rate depends on the perturbation amplitude. This result is consistent with the predictions of linear theory. Numerical simulations without friction show that nonlinear Parker instability can be triggered below the linear threshold of the field strength, when the difference in superadiabaticity along the tube is sufficiently large. A localised downflow acting on a toroidal tube in the overshoot region leads to instability depending on the parameters describing the flow, as well as the magnetic field strength. We determined ranges of the flow parameters for which a linearly Parker-stable magnetic flux tube is stored in the middle of the overshoot region for a period comparable to the dynamo amplification time. } {The longitudinal flow instability driven by frictional interaction of a flux tube with its surroundings is relevant to determining the storage time of magnetic flux in the solar overshoot region. The residence time for magnetic flux tubes with $2\\times 10^{21}$Mx in the convective overshoot layer is comparable to the dynamo amplification time, provided that the average speed and the duration of the downflow do not exceed about 50~m~s~$^{-1}$ and 100~days, respectively, and that the lateral extension of the flow is smaller than about $10^\\circ$. } ", "introduction": "\\label{sec:intro} Observations of large solar active regions are indicative of an organised subsurface magnetic field along the azimuthal (east-west) direction, with opposite polarity orientations (from east to west or vice versa) in the northern and southern hemispheres. Emerging magnetic structures are in a filamentary state, in the form of magnetic flux concentrations (flux tubes) of various sizes (e.g., sunspots, pores). Flux tubes rise in the convection zone, emerge at the surface, and form bipolar magnetic regions, which follow polarity rules (Hale's law) and systematic tilt angles (Joy's law). Theoretical studies indicate that weak magnetic fields are transported by flux expulsion \\citep{msch84} and convective pumping \\citep{tobias01} to the lower boundary of the convection zone, where the toroidal magnetic field is amplified by radial and latitudinal velocity shear in the solar tachocline \\citep[see e.g.,][]{sis06}. The stably stratified lower convection zone, in particular the convective overshoot layer, is a likely location for the generation and storage of the large-scale azimuthal magnetic flux. Numerical studies simulating a layer of horizontal magnetic field in the bottom of the solar convection zone indicate that an initially uniform field underlying a field-free layer leads to the formation of magnetic flux tubes by magnetic Rayleigh-Taylor instability \\citep[e.g.,][]{fan01}. Following its formation, a toroidal magnetic flux tube can reach a mechanical equilibrium close to the bottom of the convection zone \\citep{mi92}. The emergence of magnetic flux tubes driven by magnetic buoyancy and the properties of active regions (low-latitude emergence, tilt angles, proper motions of sunspots) require azimuthal flux densities of the order of $10^5$~G in the overshoot region \\citep[][]{dsilva93,fan94,cale95,msch96}. The possibilities for the generation of these flux densities have been reviewed by \\citet{schrempel02}, \\citet{mschafm03}, and \\citet{afm07}. The average duration of the solar activity cycle is about 11 years and the total magnetic flux emerging within one cycle ranges between orders of $10^{24}-10^{25}$~Mx. The amplification of the toroidal magnetic field to its maximum strength in about half an activity cycle raises the following question: how can the toroidal flux be stored stably for at least a few years in the dynamo amplification region? A related problem concerns the feedback of magnetic flux loss on the rate of toroidal flux generation. To more clearly understand magnetic flux generation and storage in the convective overshoot region, it is important to determine and constrain the effects of flows on the stability and dynamics of magnetic flux tubes. Here, we focus on: (a) the nonlinear development of flow-induced flux tube instabilities, which can be dynamically significant in the course of toroidal field amplification; and (b) effects of perpendicular flows on the storage of toroidal flux tubes. Consequently, some of the flow properties prevailing in the overshoot region can also be constrained, by requiring that flux tubes with field strengths of up to a few times $10^4$~G are stored in the overshoot region for about a few years. We consider the possibility that the friction-induced instability \\citep[][]{pap2,pap3} leads to flux loss from the overshoot layer for field strengths below the Parker instability limit. We refer to such Parker-stable flux tubes as ``sub-critical'' throughout the paper. We investigate effects of flows on magnetic flux tubes in mechanical equilibrium to obtain quantitative estimates, and to answer the following question: how strongly do ($i$) finite-amplitude perturbations of flux tubes by perpendicular flows \\citep[see also][paper~I]{pap1} and ($ii$) the frictional interaction of the tube with its surroundings limit the residence time of flux tubes in the overshoot region? We approach the problem by considering the nonlinearity of friction-induced instability (Sect.~\\ref{sec:friction}), including a discussion of the effects of finite perturbations on the magnetic buoyancy instability (Sect.~\\ref{ssec:delta}), and determine the ranges of flow parameters that allow the storage of sub-critical flux tubes subject to radial flows (Sect.~\\ref{sec:flows}). The analyses and simulations were carried out in the framework of the thin flux tube approximation \\citep{spruit81}, which at present is the only existing approach that can deal with the small magnetically induced variations in density, pressure, and temperature corresponding to the high plasma $\\beta$ ($>10^5$) of the deep solar convection zone. ", "conclusions": " \\begin{itemize} \\item The flow instability driven by the frictional coupling of transversal MHD waves with external flows is nonlinear in the sense that the growth rate is a function of the initial perturbation amplitude. This is consistent with the results of the linear stability analysis of paper~III, in which the parameter $\\alpha$ is proportional to the speed of the perpendicular flow and thus to the amplitude of the subsequent perturbation. Therefore, the perturbation amplitude determines the $\\alpha$ parameter in the linear analysis. \\item Significant buoyancy variations ($\\delta$-effect) along magnetic flux tubes can lead to a nonlinear buoyancy instability. The $\\delta$-effect most likely occurs in regions of large radial gradients of superadiabaticity such as the bottom of the overshoot layer, where a sufficient $\\delta$ variation along the tube ($\\Delta\\delta\\gtrsim 10^{-6}$) is easier to attain than in the upper layers. \\item For flux tubes in the solar overshoot layer, we have established links between the perpendicular flow velocity amplitude, its spatial (azimuthal) extension, and the resulting radial displacement and instabilities induced by $\\delta$-effect or longitudinal flows. \\item To store magnetic flux with flux densities between $10^4$-$10^5$~G for times comparable to the dynamo amplification time in the convective overshoot layer, the average flow speed and the flow duration must not exceed about 50~m~s~$^{-1}$ and 100~days, respectively, and that the azimuthal extension of the flow is lower than about $10^\\circ$. If we assume that these conditions prevail in the solar overshoot region, then magnetic fluxes of up to $10^{22}$~Mx can be stored within thin flux tubes during the dynamo amplification phase. \\end{itemize}" }, "0910/0910.3296_arXiv.txt": { "abstract": "}[2]{{\\footnotesize\\begin{center}ABSTRACT\\end{center} \\vspace{1mm}\\par#1\\par \\noindent {~}{\\it #2}}} \\newcommand{\\TabCap}[2]{\\begin{center}\\parbox[t]{#1}{\\begin{center} \\small {\\spaceskip 2pt plus 1pt minus 1pt T a b l e} \\refstepcounter{table}\\thetable \\\\[2mm] \\footnotesize #2 \\end{center}}\\end{center}} \\newcommand{\\TableSep}[2]{\\begin{table}[p]\\vspace{#1} \\TabCap{#2}\\end{table}} \\newcommand{\\FigCap}[1]{\\footnotesize\\par\\noindent Fig.\\ % \\refstepcounter{figure}\\thefigure. #1\\par} \\newcommand{\\TableFont}{\\footnotesize} \\newcommand{\\TableFontIt}{\\ttit} \\newcommand{\\SetTableFont}[1]{\\renewcommand{\\TableFont}{#1}} \\newcommand{\\MakeTable}[4]{\\begin{table}[htb]\\TabCap{#2}{#3} \\begin{center} \\TableFont \\begin{tabular}{#1} #4 \\end{tabular}\\end{center}\\end{table}} \\newcommand{\\MakeTableSep}[4]{\\begin{table}[p]\\TabCap{#2}{#3} \\begin{center} \\TableFont \\begin{tabular}{#1} #4 \\end{tabular}\\end{center}\\end{table}} \\newenvironment{references}% { \\footnotesize \\frenchspacing \\renewcommand{\\thesection}{} \\renewcommand{\\in}{{\\rm in }} \\renewcommand{\\AA}{Astron.\\ Astrophys.} \\newcommand{\\AAS}{Astron.~Astrophys.~Suppl.~Ser.} \\newcommand{\\ApJ}{Astrophys.\\ J.} \\newcommand{\\ApJS}{Astrophys.\\ J.~Suppl.~Ser.} \\newcommand{\\ApJL}{Astrophys.\\ J.~Letters} \\newcommand{\\AJ}{Astron.\\ J.} \\newcommand{\\IBVS}{IBVS} \\newcommand{\\PASP}{P.A.S.P.} \\newcommand{\\Acta}{Acta Astron.} \\newcommand{\\MNRAS}{MNRAS} \\renewcommand{\\and}{{\\rm and }} { We report time-series photometry for 16 variable stars located in the central part of the globular cluster NGC~6752. The sample includes 13 newly identified objects. The precision of our differential photometry ranges from 1~mmag at $V=14.0$ to 10~mmag at $V=18.0$. We detected 4 low amplitude variables located on the extended horizontal branch (EHB) of the cluster. They are candidate binary stars harboring sdB subdwarfs. A candidate degenerate binary was detected about 2 mag below the faint end of the EHB. The star is blue and its light curve is modulated with a period of 0.47~d. We argue that some of the identified variable red/blue stragglers are ellipsoidal binaries harboring degenerate stars. They have low amplitude sine-like light curves and periods from a few hours to a few days. Spectroscopic observations of such objects may lead to the detection of detached inactive binaries harboring stellar mass black holes or neutron stars. No binaries of this kind are known so far in globular clusters although their existence is expected based on the common occurrence of accreting LMXBs and millisecond pulsars. An eclipsing SB1 type binary was identified on the upper main sequence of the cluster. We detected variability of optical counterparts to two X-ray sources located in the core region of NGC~6752. The already known cataclysmic variable B1=CX4 experienced a dwarf nova type outburst. The light curve of an optical counterpart to the X-ray source CX19 exhibited modulation with a period 0.113~d. The same periodicy was detected in the HST-ACS data. The variable is located on the upper main sequence of the cluster. It is an excellent candidate for a close degenerate binary observed in quiescence. } {\\bf Key words:} {\\it stars: dwarf novae - novae, cataclysmic variables -- globular clusters: individual: NGC 6752 -- stars: horizontal-branch -- binaries: eclipsing} ", "introduction": "\\label{sect:intro} NGC~6752 is a nearby globular cluster whose low reddening and relatively high Galactic latitude ($(m-M)_{V}=13.02$, $E(B-V)=0.04$, $b=-25.6$~deg; Harris 1996) make it an attractive target for detailed studies. Optical counterparts for 12 out of 19 faint X-ray sources detected with $Chandra$ were reported by Pooley et al. (2002). This sample included ten likely cataclysmic variables (CVs) and 1-3 RS~CVn or BY Dra stars. Two of the candidate CVs were detected and studied earlier by Bailyn et al. (1996) based on HST/WFPC2 data. So far there have been no reports of dwarf novae type outbursts in the field of NGC~6752. The cluster is known to host five millisecond pulsars (D'Amico et al. 2002), one of which has an optical counterpart (Ferraro et al. 2003; Bassa et al. 2003). A wide-field CCD based survey of NGC~6752 conducted by Thomson et al. (1999) led to the detection of eleven photometric variables, seven of which were classified as contact binaries, and three as SX~Phe stars. The issue of membership status of these variables remains open. In particular, Rucinski (2000) argues that the group of contact binaries is dominated by field interlopers. The cluster has a rich population of blue horizontal branch (BHB) and extreme horizontal branch (EHB) stars (Buonanno et al. 1986; Momany et al, 2002). Until recently not a single photometric variable was known among the horizontal branch stars in NGC~6752. Catelan et al. (2008) reported possible detection of a BHB pulsator, while Kaluzny \\& Thompson (2008) detected four variables located on the BHB/EHB of the cluster. Only one spectroscopic binary was found among a few dozen BHB/EHB stars observed by Moni Bidin et al. (2008). This result was unexpected, as close binaries are common among field sdB stars (Maxted et al. 2001; Napiwotzki et al. 2004). The photometric survey presented here was conducted as a part of the CASE project (Kaluzny et al. 2005). It is complementary to the wide-field study presented by Thompson et al. (1999). The new data, collected with a larger telescope at a finer xpatial scale, allowed a more detailed study of the central part of the cluster. Since the pioneering work by Mateo et al. (1990) several globular clusters have been surveyed with CCD photometry for faint variables, and in particular for main sequence binaries. These surveys, usually conducted with 1-m class telescopes, are very incomplete in the cluster core regions, arising mainly from the saturation of stellar profiles of densely packed bright stars. This saturation is hard to avoid as exposures times have to be at least a few minutes long to assure sufficient S/N for the relatively faint main-sequence stars. The problem can be partially overcome by using larger telescopes with finer pixel scales producing images with a reduced number of saturated stars. An even better solution is to use imaging capabilities of the $HST$, but so far only one cluster has been systematically surveyed for variability with this instrument (47 Tuc, Gilliland et al. 2000; Albrow et al. 2001). Results published for a few other clusters observed with the $HST$ are based on fragmentary data. It is expected that most of the eclipsing variables are located in the central parts of globular clusters. First, by definition, roughly 50\\% of all stars are observed within the half-light radius of a given cluster. Second, dynamical considerations indicate that most of the binary stars migrate toward the core regions due to mass segregation (Stodolkiewicz 1986; Hut 1992). As we show below, our adopted observing strategy along with careful data reduction has allowed a successful search for variability in a large sample of stars in the central part of NGC~6752. Good quality photometry extends down to $V\\approx 21$~mag and the light curves of the brightest stars show an rms of about 1~mmag. ", "conclusions": "\\label{sect:summary} We have obtained time series $BV$ photometry for about 40 thousand stars from the central area of the globular cluster NGC~6752. The observing strategy and careful reduction of the data resulted in a photometric precision of 4 mmag stars with $V<16.0$ and 1-2 mmag at $V<15$, rising to 0.01 mag at $V\\approx 18.5$. The sample included 72 hot subdwarf candidates. No pulsation variability was detected for any of them. We have detected, however, four low-amplitude variables which may be binary EHB/BHB stars. Spectroscopic follow up is needed to reveal their actual nature. Preliminary periods were established for two of them. A faint blue variable with $V\\approx 20.7$ was also detected. Its light curve is likely periodic with $P\\approx 0.47$~d. The variable is a good candidate for a degenerate binary belonging to the cluster. One of four periodically variable blue stragglers is a likely $\\gamma$~Dor star. If confirmed, it would be the first variable of this type detected in globular clusters. We detected three variable yellow/red stragglers with $15.0 10^6$ times the mass of the black hole in Cyg X-1 (like most AGN in our sample). This suggests that each one of the AGN points in Fig.~3 corresponds to just one of the Cyg X-1 points plotted in the same figure. We found that the AGN and Cyg X-1 ``$\\Gamma - \\nu_{\\rm norm}$\" relations are similar but not the same. In Section 5 we discussed briefly some implications of this result. In the paragraphs below we discuss the implications of our results in more detail. The main aim of the discussion below is to investigate the constrains that the AGN spectral-timing relation, and its comparison with the similar relation in Cyg X-1, impose on thermal Comptonisation models, based on the particular assumption that both the spectral and timing properties of accreting systems are driven by accretion rates variations. We point out that other interpretations are also possible; see for example Kylafis \\etal\\ (2008) for an alternative explanation of the Cyg X-1 spectral-timing relation, which does not assume that X--rays are produced by thermal Comptonisation. We also point out that, given the small number of objects in our sample, and the unavoidable uncertainty in the derived parameters of the AGN spectral-timing relation, the values of the various parameters in the equations below are somehow uncertain. As we showed in Section 6.1, the AGN spectral-timing relation and the M06 results imply that $\\Gamma_{\\rm AGN} \\approx 2.3 \\dot{m}_{\\rm E}^{0.07}$. If X--rays in AGN are produced by thermal Comptonisation, we expect\\footnote{The discussion in these paragraphs is based on the assumption of an outflowing corona, hence we adopt the results of Malzac \\etal, 2001. Similar conclusions can also be drawn if we assume a static corona.} that $ \\Gamma_{\\rm AGN} = 2.2 (L_{\\rm s}/L_{\\rm diss})^{0.07}$. Therefore, the observations [i.e. $\\Gamma_{\\rm AGN} \\approx 2.3 \\dot{m}_{\\rm E}^{0.07}$] are consistent with the thermal Comptonisation model predictions [$ \\Gamma_{\\rm AGN} = 2.2 (L_{\\rm s}/L_{\\rm diss})^{0.07}$], only if \\begin{equation} (L_{\\rm s}/L_{\\rm diss})\\approx 2 \\dot{m}_{\\rm E}. \\end{equation} \\noindent An obvious implication of this result is that, if a certain fraction, say $f_{\\rm s}$, of the total accretion power, $P_{\\rm tot}$ $(=\\eta\\dot{m}_{\\rm E}c^2,$ where $\\eta$ is the efficiency of the accretion process), is converted to disc luminosity (i.e. $L_{\\rm s}=f_{\\rm s}P_{\\rm tot}$), while $L_{\\rm diss}=f_{\\rm diss}P_{\\rm tot}$, then the ratio $(f_{\\rm s}/f_{\\rm diss})$ should not remain constant for a given object, but should rather increase with increasing accretion rate. Suppose that X--rays from Cyg X-1 in its low/hard state (LH) are also produced by thermal Comptonisation. In this case, thermal Comptonisation models predict that $\\Gamma_{\\rm CygX-1} = 2.2 (L_{\\rm s}/L_{\\rm diss})^{0.13}$, or \\begin{equation} \\Gamma_{\\rm CygX-1}\\approx 2.4 \\dot{m}_{\\rm E}^{0.13}, \\end{equation} \\noindent if we accept that equation (1) holds in this case as well. According to K\\\"{o}rding \\etal\\ (2007), the normalisation of the $\\nu_{\\rm norm}$ vs $\\dot{m}_{\\rm E}$ relation for the hard-state GBHs is $\\sim 8$ times smaller than the normalisation in the case of the AGN in our sample. So, if $\\nu_{\\rm norm, Cyg X-1/LS}\\approx 375\\dot{m}_{\\rm E}$, as opposed to $\\nu_{\\rm norm, AGN}\\approx 3000\\dot{m}_{\\rm E}$, then $\\dot{m}_{\\rm E}\\approx 3\\times 10^{-3}\\nu_{\\rm norm, Cyg X-1/LS}.$ We can now substitute $\\dot{m}_{\\rm E}$ in equation (2) to determine the $\\Gamma -\\nu_{\\rm norm}$ relation for Cyg X-1 in LH state: \\begin{equation} \\Gamma_{\\rm CygX-1/LS}\\approx 1.1 \\nu_{\\rm norm}^{0.13}. \\end{equation} \\noindent The dashed line in the top panel of Fig.~3 indicates this relation. The agreement between the {\\it predicted} $\\Gamma-\\nu_{\\rm norm}$ relation and the Cyg X-1 data is rather good up to $\\nu_{\\rm norm}\\sim 70$. The discussion so far suggests the following picture: a) X-rays from the AGN we studied are produced by thermal Comptonisation. The $\\Gamma-\\nu_{\\rm norm}$ relation we observe is consistent with the predictions of thermal Comptonisation models but only if the $(L_{\\rm s}/L_{\\rm diss})$ ratio, and hence the $(f_{\\rm s}/f_{\\rm diss})$ ratio as well, increase proportionally with accretion rate. b) X-rays in Cyg X-1 are also produced by thermal Comptonisation. Taking into account the fact that the normalisation of the $\\nu_{\\rm norm}-\\dot{m}_{\\rm E}$ relation is $\\sim 8$ times smaller in Cyg X-1 than in the AGN in our sample (K\\\"ording \\etal\\ 2007), the {\\it predicted} $\\Gamma-\\nu_{\\rm norm}$ relation agrees well with the Cyg X-1 data up to $\\nu_{\\rm norm}\\approx 70$. c) In the case when $\\Gamma_{\\rm AGN}=\\Gamma_{\\rm Cyg X-1}$, we expect that ($L_{\\rm s}/L_{\\rm diss})_{\\rm Cyg X-1}^{0.13}=(L_{\\rm s}/L_{\\rm diss})_{\\rm AGN}^{0.07}$, and $\\dot{m}_{\\rm AGN,E}^{0.07}=\\dot{m}_{\\rm Cyg X-1,E}^{0.13}$ (using equation 1). Consequently, $\\dot{m}_{\\rm E,AGN}=\\dot{m}_{\\rm E,Cyg X-1}^{1.9}$ and, since $\\dot{m}_{\\rm E}<1$, AGN should operate on a lower accretion rate than Cyg X-1 when the spectral slope is the same in both systems. Furthermore, since $\\Gamma_{\\rm AGN}\\approx 1.3 \\nu_{\\rm norm,AGN}^{0.07}$ and $\\Gamma_{\\rm Cyg X-1}\\approx 1.1 \\nu_{\\rm norm,Cyg X-1}^{0.13}$ (when $\\Gamma\\la 2.1-2.2$), then $\\Gamma_{\\rm AGN}=\\Gamma_{\\rm Cyg X-1}$ implies that $1.3 \\nu_{\\rm norm,AGN}^{0.07}=1.1 \\nu_{\\rm norm,Cyg X-1}^{0.13}$, and $\\nu_{\\rm norm,Cyg X-1}\\approx 3.5\\nu_{\\rm norm,AGN}^{0.55}$. Therefore, when the spectral slope is the same (and less than $\\sim 2.1-2.2$) in AGN and Cyg X-1, the former should operate at a lower accretion rate but their characteristic time scales should be shorter than those in Cyg X-1 (when normalised to the respective BH mass), because the normalisation of the AGN $\\dot{m}-\\nu_{\\rm norm}$ relation is significantly larger than the normalisation of the respective Cyg X-1 relation in LH state. The vertical dot-dashed line in the top panel of Fig.~3 indicates the value $\\nu_{\\rm norm}=70$. For Cyg X-1, this normalised frequency corresponds to $\\nu_{1}=1$ Hz and $\\nu_2=5$ Hz for the centroid frequency of the low and higher frequency Lorentzians, respectively. As $\\nu_{\\rm norm}$ (i.e. the accretion rate) increases even more, then i) the ratio $\\nu_{1}/\\nu_2$ increases as well (see Fig.~2 in A06), ii) the contribution of the Lorentzians to the root mean square variability amplitude decreases, and iii) the ``bending\" power-law component in the PSD appears and its contribution to the source variability amplitude increases (see Fig.~7 in A06), i.e. the Cyg X-1 power spectrum changes from a hard to a soft-state shape. At $\\nu_{\\rm norm}\\approx 70$, the average {\\it observed} spectral slope of Cyg X-1 is $\\sim 2.1$ (see Fig.~3), while at higher frequencies the spectral slope is steeper. Consequently, the region defined by $\\Gamma<2.1$ and $\\nu_ {\\rm norm}<70$ corresponds to the ``hard state region\" for Cyg X-1 in the ``spectral-timing plane\" shown in Fig.~3. The discrepancy between the predicted spectral-timing relation and the Cyg X-1 data when $\\Gamma\\ma 2.1$ and $\\nu_{\\rm norm}\\ma 70$ cannot be explained by the fact that the normalisation of the $\\nu_{\\rm norm} - \\dot{m}_{\\rm E}$ relation increases by a factor of $\\sim 8$ in the high/soft state. If that were the case, the spectral-timing relation in this state should be similar to the one defined by equation (3), but with a {\\it smaller} normalisation (opposite to what we observe). One possibility is that the $(L_{\\rm s}/L_{\\rm diss})-\\dot{m}_{E}$ relation (equation 1) changes in the high/soft state. However, in this case we would have to accept that the X--ray source does {\\it not} operate in the same way in AGN and Cyg X-1: when $\\nu_{\\rm norm}>70$, both Cyg X-1 and the AGN in our sample follow the same $\\nu_{\\rm norm} - \\dot{m}_{\\rm E}$ relation (M06, K\\\"ording \\etal\\ 2007). Therefore, as long as $\\nu_{\\rm norm}>70$, a given $\\nu_{\\rm norm}$ value implies the same accretion rate in both systems. The fact that the AGN spectral-timing relation is valid up to $\\nu_{\\rm norm}\\approx 1000$ should then imply that, for the same accretion rate, the ``$(L_{\\rm s}/L_{\\rm diss}) - $ accretion rate\" relation is different in AGN and Cyg X-1. In Cyg X-1, $\\Gamma\\approx 2.1$ implies that $2.2 (L_{\\rm s}/L_{\\rm diss})^{0.13}\\approx 2.1$, and hence $(L_{\\rm s}/L_{\\rm diss})\\approx 0.7$. Another possible explanation then for the discrepancy between the Cyg X-1 data and the predicted spectral-timing relation above $\\nu_{\\rm norm}=70$ is the following: equation (1) holds until $(L_{\\rm s}/L_{\\rm diss})\\approx 0.7$, at which point the hot corona is significantly cooled down, and the thermal X--ray emission component is weak. It is possible then that at high accretion rates a separate, possibly non-thermal, X--ray component emerges, and dominates the X-ray emission in the soft state. If that is the case, the $\\Gamma- L_{\\rm s}/L_{\\rm diss}$ relation we have assumed above is not valid, hence the predicted spectral-timing relation does not fit the data any more. Even if the picture drawn above is correct, there are important issues regarding the relation between AGN and GBH states which still remain unresolved. In particular, the answer to the question whether the AGN in our sample are ``soft\" or ``hard\" state systems is far from clear. There are indications that they are analogous to Cyg X-1 in its soft state. For example, the radio emission in Cyg X-1 in its LH state is enhanced. On the other hand, Panessa \\etal\\ (2007) have shown that, for the same $L_{\\rm X}/L_{\\rm E}$ ratio, the radio luminosity of Seyfert galaxies is $\\sim 8-10$ times lower than the radio luminosity of hard-state GBHs, even when the BH mass difference is properly taken into account (7 of the objects in our sample are also included in their sample). Furthermore, the AGN in our sample follow the soft-state ``characteristic time scale -- accretion relation\" in GBHs, and a ``bending\" power-law is the dominant component in the power spectrum of those objects with good enough light curves to accurately study their PSD (e.g. NGC~4051, M$^{\\rm c}$Hardy \\etal\\ 2004; NGC~3227 and NGC~5506, UM05; MCG~6-30-15, M$^{\\rm c}$Hardy \\etal\\ 2005; and perhaps NGC~3783, Summons \\etal\\ 2007), implying a soft-state like PSD in these objects. High quality light curves for low accretion rate AGN are necessary to investigate whether any AGN with hard-state like power spectra exist or not. However, although the radio emission strength and the timing properties of many objects in our sample are soft-state like, their spectral properties are {\\it not}, as the average spectral slope is smaller than $2.1$ in most cases. If indeed the {\\it spectral} hard-to-soft state transition corresponds to $(L_{\\rm s}/L_{\\rm diss})\\sim 0.7$, at which point the thermal corona emission is weak, and a different component dominates the X--ray emission, then we should expect this transition to happen when $\\Gamma\\ma 2.15$ in AGN. Given the AGN $\\Gamma-\\nu_{\\rm norm}$ relation, this slope corresponds to $\\nu_{\\rm norm}\\approx 2500$. At even higher normalised frequencies, we should then expect the AGN spectral-timing relation to break (like in Cyg X-1 at $\\nu_{\\rm norm}\\approx 70$). Obviously, more data are necessary to confirm that this is indeed the case in AGN. The data so far suggest that while in AGN the {\\it timing} properties transition from hard-to-soft state happens at least as low as $\\nu_{\\rm norm}\\approx 100$, the {\\it spectral} properties transition should happen at much higher accretion rates. This is opposite to what we observe in Cyg X-1, where both the spectral and timing properties change from the hard to soft state, at the same accretion rate (indicated by the value $\\nu_{\\rm norm}\\approx 70)$. Perhaps the timing properties in accreting objects are determined by accretion disc variations only, and the hard-to-soft state transition materialises at a certain accretion rate, irrespective of whether the soft disc luminosity is strong enough for $(L_{\\rm s}/L_{\\rm diss})> 0.7$, i.e. strong enough to cool down the hot corona. In this case, we would expect the AGN hard-to-soft timing properties transition to appear at $\\nu_{\\rm norm}\\approx 70$ as well. Due to the cooler disc temperature though, the AGN spectral soft-to-hard state transition happens at higher accretion rates (i.e. at a higher $\\nu_{\\rm norm}$ value in the spectral-timing plane of Fig.~3). Further progress in understanding the relation between AGN and GBHs can be made when we know how the accretion rate determines the characteristic frequency in accreting compact objects (assuming that it is just the accretion rate that determines $\\nu_{\\rm norm}$ in these systems)." }, "0910/0910.2545_arXiv.txt": { "abstract": "Blue stragglers have been found in all populations. These objects are important in both stellar evolution and stellar population synthesis. Much evidence shows that blue stragglers are relevant to primordial binaries. Here, we summarize the links of binary evolution and blue stragglers, describe the characteristics of blue stragglers from different binary evolutionary channels, and show their consequences for binary population synthesis, such as for the integrated spectral energy distribution, the colour-magnitude diagram, the specific frequency, and the influences on colours etc.. ", "introduction": "Blue stragglers (BSs) are an important population component in stellar evolution as well as in star clusters. These objects have remained on the main sequence for a time exceeding that expected from standard stellar evolution theory, and they may affect the integrated spectra of their host clusters by contributing excess spectral energy in the blue and UV bands. Many mechanisms, including single star models and binary models, have been presented to account for the existence of BSs (see the review of \\cite{str93}). At present, it is widely believed that more than one mechanism plays a role for the produce of BSs in one cluster and that binaries are important or even dominant for the production of BSs in open clusters and in the field (\\cite{lan07,dal08,sol08}). Binaries may produce BSs by way of mass transfer, coalescence of the two components, binary-binary collision and binary-single star collision. The collision of binary-binary or binary-single may lead binaries to be tighter or farther apart, and it is relevant to dynamics and environment in the host cluster. In this contribution, we are only concerned with BSs resulting from the evolutionary effect of primordial binaries, i.e. mass transfer and coalescence of two components. ", "conclusions": "In this contribution, we summarized the linkage of binary evolution and BSs, and BS characteristics. As well, we showed binary population synthesis results of BSs from primordial binary evolution, such as the distribution on CMD, the contribution to ISED, the specific frequency and the influences on colours. This work was in part supported by the Chinese National Science Foundation (Grant Nos. 10603013 and 10973036,10821061 and 2007CB815406) and Yunnan National Science Foundation (Grant No. 08YJ041001)." }, "0910/0910.0589_arXiv.txt": { "abstract": "We show that an object classified as a galaxy in on-line data bases and revealed on sky survey images as a distant ring galaxy is a rare case of polar ring galaxy where the ring is only slightly inclined to the equatorial plane of the central body. Imaging information from the Sloan Digital Sky Survey (SDSS) indicates that the diameter of the ring is about 36 kpc. The SDSS data was combined with long-slit spectroscopic observations and with Fabry-P\\'{e}rot Interferometer H$\\beta$ mapping obtained at the Russian Academy of Sciences 6-m telescope. We derived the complex morphologies of this presumed ring galaxy from a combination of SDSS images and from the kinematical behaviour of the central body and of the ring, and determined the stellar population compositions of the two components from the SDSS colours, from the spectroscopy, and from models of evolutionary stellar synthesis. % The metallicity of the ring material is slightly under-abundant. The total luminosity and the total mass of the system are not extreme, but the rather high M/L$\\simeq$20 indicates the presence of large amounts of dark matter. We propose two alternative explanations of the appearance of this object. One is a ring formed by two semi-circular and tight spiral arms at the end of a central bar. The apparent inclination between the ring and the central body, and a strange kink at the North-East end of the ring, could be the result of a warp or of precession of the ring material. The object could, therefore, be explained as an extreme SBa(R) galaxy. The other possibility is that we observe a Polar Ring Galaxy where the inner object is an S0 and the ring is significantly more luminous than the central object. The compound object would then be similar to the NGC 4650A galaxy, but then it would be a rare object, with a polar component only modestly inclined to the equatorial plane of the central body. Arguments for (and against) both explanations are given and discussed, with the second alternative being more acceptable. ", "introduction": "\\label{txt:Obs_and_Red} \\subsection{SDSS data} \\label{txt:sdss} The SDSS \\citep[][]{York2000,DR5} imaging data were used to provide morphological parameters for the galaxy, colours of the central object, and colours and colour distributions for the ring. Since the source extraction routines of SDSS are not optimized for the detection and measurement of very faint surface brightness extended objects, SDSS provides only the photometry for the central component and also for those ``shreds'' of ring identified as ``stars'' by the SDSS detection software. Therefore, specialized photometry in the SDSS bands was applied to measure the properties of the central component and of the ring using the low surface brightness detection methods developed and described in \\citet{LSB_phot}. Specifically, a deep image of the galaxy and of the ring was created by combining the $g$, $r$, and $i$ images with weights, as explained in \\citet{LSB_phot}. While no physical meaning can easily be attached to photometry using such a wide and ill-defined photometric band, the resulting image is significantly deeper than the individual SDSS images and is useful to investigate the object morphology. The deep $gri$ combined image is shown in Figure~\\ref{Fig:RG1_aper}. % An outermost isophote was used to trace automatically an irregularly-shaped aperture that collects all the light from even very faint surface brightness parts of the galaxy. % At this point, a few remarks about the appearance of RG1 are in order. Figure~\\ref{Fig:RG1_aper} shows a relatively bright, elliptical-shaped central body encircled by an $\\sim$25-arcsec elliptical ring. Two roundish and relatively compact non-stellar objects are visible near the ring; one is to the East (galaxy E) and the other is in the opposite side (galaxy W), slightly more distant from the centre and fainter than galaxy E. These bodies are reminiscent of the proposed ``impactor'' galaxy in a scenario for the formation of collisional ring galaxies. In addition, the ring shows a loop-like extension or kink to the North-East. In order to perform integral and surface brightness photometry we used concentric circular annuli centered on the galaxy's photo-centre \\citep{LSB_phot}, with radii increasing outward by one SDSS pixel. The average surface brightness and the average colour were determined in each annulus. \\subsection{Long-slit spectroscopy with the 6m BTA telescope of the RAS } \\label{txt:long_slit_obs} First spectroscopic data were obtained % with the 6-meter Large Altazimuthal Telescope (BTA) of the Special Astrophysical Observatory of Russian Academy of Science (SAO RAS) on February 16/17 2001 (see Table~\\ref{t:Obs} for details). The 2001 observations used the Long-Slit spectrograph UAGS (Afanasiev et al. 1995) at the telescope prime focus equipped with a Photometrics CCD detector with 1024$\\times$1024 pixels each $24\\times24\\, \\mu m^2$ in size. Two 130\\arcsec\\ long-slit spectra were obtained in blue and in red with a grating having 651 grooves/mm yielding a dispersion of 2.4~\\AA/pixel and a spectral resolution of 6--7~\\AA\\,. The slit position (see Fig.~\\ref{Fig:RG1_aper}) was chosen to cross the central bright region close to the direction of the major axis of the inclined ring. The wavelength ranges of the obtained spectra are given in Table~\\ref{t:Obs}. A 2\\arcsec\\ wide slit was used in both cases. The scale along the slit was 0.39\\arcsec/pixel, very similar to that of the SDSS images. Reference spectra of an Ar--Ne--He lamp were recorded before or after each observation to provide wavelength calibration. Spectrophotometric standard stars from Bohlin (1996) were observed for flux calibration. New spectral observations with the BTA were obtained by AM on January 15/16 and on February 12/13, 2008, as also detailed in Table~\\ref{t:Obs}. These were collected using the long-slit mode of the SCORPIO universal focal reducer (Zasov et al. 2008). A volume-phase holographic grating with 1200 grooves mm$^{-1}$ was used, providing a dispersion of 0.84 \\AA\\,pixel$^{-1}$ and yielding spectra that covered the range $\\sim$5600\\AA\\, to $\\sim$7310\\AA. The spectral resolution of these spectra was $\\sim$5 \\AA\\, and the spatial sampling was 0.36 arcsec pixel$^{-1}$. The locations of the entrance slit for the two observed positions are shown in Figure~\\ref{Fig:RG1_aper}. The primary data reduction included cosmic-ray removal with the {\\tt MIDAS}\\footnote{{\\tt MIDAS} is an acronym for the European Southern Observatory package --- Munich Image Data Analysis System.} software package, and bias subtraction and flat-field correction with the {\\tt IRAF}\\footnote{{\\tt IRAF}: the Image Reduction and Analysis Facility is distributed by the National Optical Astronomy Observatories, which is operated by the Association of Universities for Research in Astronomy, In. (AURA) under cooperative agreement with the National Science Foundation (NSF).} software package. For subsequent reduction of the long-slit spectra we used {\\tt IRAF}. In case of UAGS data, after wavelength mapping and night sky subtraction, each 2D frame was corrected for atmospheric extinction and was flux calibrated. 2D spectra of different spectral regions were combined after that stage. RG1 has an SDSS spectrum in the public archives. Since this was obtained with a single fiber, it samples only the central three arcsec of the central body. We extracted the innermost three arcsec spectrum from our long-slit reduced UAGS 2D spectrum and compared it with the SDSS spectrum. Although the comparison shows that the two spectra match within 10\\%, we assumed that the SDSS spectrum has a more accurate flux distribution and constructed the correction function based on it. The reason for this assumption is that SDSS is observing simultaneously with 640 fibers, where some are targeted on flux standards. This implies that each SDSS spectroscopic measurement has a few standard stars out of the 640 spectra collected and their fluxes are used for calibration. For 6m blue and red parts of the final spectrum where observed separately in not-so-perfect weather conditions. Additionally, blue and red parts of the flux standards were observed separately. Thus, from our point of view, ``many and simultaneous'' calibrations are better than ``few and separate'' and we selected the SDSS spectrum to derive the correction function. The derived function was applied to the reduced UAGS 2D spectrum and the result is shown in Figure~\\ref{Fig:2D-spectrum} together with the 1D spectra of the central component and the outer part of the ring. To obtain the line-of-sight velocity distribution along the slit we used a classical cross-correlation method described in detail by \\citet{Zasov00}. All emission lines were measured with the MIDAS programs described in detail in \\citet{SHOC,Sextans}. \\begin{figure*} {\\centering \\includegraphics[clip=,angle=-90,width=17.0cm]{fig2a.eps} \\includegraphics[clip=,angle=-90,width=8.0cm]{fig2b.eps} \\includegraphics[clip=,angle=-90,width=8.0cm]{fig2c.eps} } \\caption{{\\it Upper panel:} Part of the reduced BTA UAGS 2D spectrum of RG1 for $\\rm PA \\sim 58^{\\circ}$. NE is up. The two arcsec wide slit was positioned on the central component and along the major axis. The ring region exhibits the redshifted [\\ion{S}{ii}], H$\\alpha$, [\\ion{N}{ii}], [\\ion{O}{iii}], H$\\beta$, H$\\gamma$, and [\\ion{O}{ii}] emission lines with measurable intensities on a visible, though faint, continuum that seems stronger on the blue side of the image. The two Ca II H \\& K lines and the CH G-band absorption are visible near the blue end of the central object spectrum. The spectrum of the central object and of both sides of the ring is easily visible for a distance of $\\pm$14\\arcsec\\ along the slit. At the adopted distance of 250 Mpc 1\\arcsec\\ = 1.2 kpc and the image vertical extent is $\\sim$33.6 kpc. Horizontal lines to the right of the panel show parts of 2D spectrum, which were averaged and used in Section~\\ref{txt:abund}. {\\it Lower panel:} The 1D spectra extracted from the 2D spectrum observed at $\\rm PA \\sim 58^{\\circ}$: the central component of RG1 (left, region {\\it c}) and outer part of the ring (right, region {\\it 2}). Some detected emission and absorption lines have been marked. The spectrum of the central component is equivalent with the one in the SDSS database. \\label{Fig:2D-spectrum}} \\end{figure*} \\subsection{Observations with the scanning Fabry-P\\'{e}rot interferometer} The UAGS spectra showed that the H$\\beta$ line from the ring is fairly well-defined and not confused with other spectral features. It is therefore beneficial to attempt a two-dimensional kinematic mapping with field spectroscopy. Observations were performed during the night of 26-27 October 2008 with the scanning Fabry--P\\'{e}rot Interferometer (FPI) mounted within the SCORPIO focal reducer (Afanasiev \\& Moiseev 2005) at the prime focus of the SAO RAS 6-m telescope. The desired spectral interval in the neighbourhood of the redshifted H$\\beta$ line was selected by a narrow-band filter with full-width at half maximum (FWHM) of 21\\AA\\,. The free spectral interval between the neighbouring interference orders was 18\\,\\AA\\, ($\\sim$1100 km/s). The FPI resolution, defined as the $FWHM$ of the instrumental profile, was $1.5$\\AA\\, ($\\sim90$ km/s) for a 0.55\\AA\\, sampling. The detector was a $2048\\times2048$ EEV 42-40 CCD operating in on-chip binned $4\\times4$ pixel mode to reduce readout time and match the resultant pixel to the seeing size. The final image scale and field of view were 0\\farcs7/pixel and 6\\farcm1$\\times$6\\farcm1, respectively. We collected 32 successive interferograms of the object with different spacings of the FPI plates. The total exposure was 96~min and the seeing during the observations varied from 1\\farcs5 to 2\\farcs0. The observations were reduced using an IDL-based software package described by Moiseev (2002) and Moiseev \\& Egorov (2008). Following the primary reduction, night-sky line subtraction, and wavelength calibration, the frames were combined into a data cube where each pixel in the $512\\times512$ field contains a 32-channel spectrum centered on the H$\\beta$ line sampled at 0.55\\AA\\ per bin. The final angular resolution after optimal smoothing matches the 2\\farcs4 seeing. \\begin{figure*} \\centerline{ \\includegraphics[width=9cm]{Fig_3a.eps} \\includegraphics[width=9cm]{Fig_3b.eps}} \\caption{Results of FPI observations: intensity distribution of the H$\\beta$ line emission (left panel, with a linear intensity scale) and line-of-sight velocity field of the ionized gas (right panel). The coordinate origin coincides with the dynamical centre of the object.} \\label{fig:FPIimage} \\end{figure*} Figure~\\ref{fig:FPIimage} shows an image of the object in the H$\\beta$ line (left panel) obtained by Gaussian fitting the line emission in the FPI data cube, and subtracting a properly scaled combination of the line-free channels that contain the continuum contribution. The right panel of Figure~\\ref{fig:FPIimage} shows the distribution of radial velocities over the same image. Note that there is line emission in the central object and in the bridge connecting it to the ring. Note also the enhancements in the ring itself; these features shall be discussed below. Figure~\\ref{fig:FPIimage} shows also an emission patch $\\sim$10\\arcsec\\ away from the ring at (--$\\Delta\\alpha$, $\\Delta\\delta$)=(13, 13); this corresponds to a faint patch in the SDSS $gri$ deep image (see Figure~\\ref{Fig:RG1_aper}).% ", "conclusions": "\\label{txt:summ} We presented above SDSS photometry, long-slit spectroscopy, and FPI data about a possible ring galaxy. The observations have shown that: \\begin{enumerate} \\item The central body shows a S\\'{e}rsic surface brightness distribution with $n\\simeq1.2$, is much redder than the ring, and exhibits emission lines on top of an absorption spectrum. \\item The ring is blue, with strong emission lines, maintains the same colour through its width, and is connected to the CB by a linear structure. The H${\\alpha}$-to-H${\\beta}$ emission line ratio indicates A$_B$=2.7 in the ring. This, and the forbidden line ratios, indicate that the ring metallicity is 12+log(O/H)$\\sim$8.32, lower than expected considering the ring luminosity. \\item The CB stars have a velocity dispersion commensurate with the maximal ring rotation velocity. Their rotation velocity at the outermost locations joins smoothly with the ring rotation. \\item The ionized gas in the ring shows an increasing rotation curve almost to its outermost point. The systemic velocity of each ring segment decreases with galacto-centric distance. The plane of the ring is tilted by 58$^{\\circ}\\pm10^{\\circ}$, or 73$^{\\circ}\\pm11^{\\circ}$, from the plane of the CB. \\item The rotation curve indicates a total dynamical mass of 4$\\times 10^{11}$ M$_{\\odot}$ within the outer edge of the ring. Since the entire object has M$_g$=--20.35, assuming a solar absolute magnitude of 5.45 mag in the same spectral band implies a total object luminosity of $\\sim2\\times 10^{10}$ L$_{\\odot}$. Therefore RG1 has an overall M/L$\\simeq$20, rather high for galaxies and implying significant amounts of dark matter in the system. \\end{enumerate} Sil'chenko \\& Moiseev (2006) showed that NGC 7742, an Sb galaxy with a nuclear star-forming ring and lacking a bar, has two global exponential stellar disks with different scale lengths and a circumnuclear inclined gaseous disk with a radius of 300 pc. The galaxy has a global gaseous counterrotating subsystem, which they interpret as a result of a past minor merger, including the appearance of the nuclear star-forming ring lacking a global bar; the ring might be produced as a resonance feature by tidally-induced oval distortions of the global stellar disk. The connection between the CB and the ring in RG1, and the smooth match of the rotation curves of the CB (from stellar absorption lines) to that of the ring (from the emission lines), indicates that RG1 could probably be an extreme case of barred spiral where the two arms form a complete circle. However, the kinematics described here, in particular the second solution obtained for the inclination, would indicate an extreme warp of the disk; while the inner regions are reasonably flat, from about 25 arcsec out to the end of the luminous body the disk turns by 60--70$^{\\circ}$ within a few arcsec. It is also possible that this is another case of polar-ring galaxy, like the well-known NGC4650A with a relatively small central stellar body and massive self-gravitating stellar and gaseous polar ring. The polar disk could have a spiral structure induced by the non-axisymmetric potential of CB; this is similar to that of a bar in a disk galaxy but without being a real bar, matching the simulations of Theis et al. (2006). The similarities with respect to NGC 4650A can be summarized as (a) a CB with a near-exponential profile, with a S\\'{e}rsic index $n\\simeq1$, (b) a bright and presumably massive, self-gravitating ring or disk, and (c) a spiral or quasi-spiral structure in the polar disk/ring. We reconsider here the issue of metal abundance in light of the possible nature of the objet. We remind the reader that the metal abundance measurement shown in Figure~\\ref{fig:line_ratios} showed a fairly constant abundance of 12+log(O/H)=8.30 to 8.65. The measurements covered ring radii 4$\\leq$R$\\leq$15 arcsec; given that the 25 mag arcsec$^{-1}$ surface brightness level is reached at R$\\simeq$12 arcsec, this implies that our abundance measurements for the ring cover normalized radii 0.4$\\leq \\frac{R}{R_{25}} \\leq$1.25. We compare this with the metallicity gradients measured by van Zee et al. (1998) for a sample of spiral galaxies, where their figure 12 shows considerable oxygen gradients over this range. In contrast to the situation regarding spiral galaxies, as shown by van Zee et al. (1998) as well as by Denicolo et al. (2001), very few PRGs have been studied to the determine their metal abundances. A few examples are UGC 5600 (Shalyapina et al. 2002) where no gradient was found, NGC 7468 (Shalyapina et al. 2004) with an inverted gradient where the outer parts are more metal-rich than the inner parts, and AM1934-563 (Brosch et al. 2007) with a normal gradient for the galaxy but a constant metallicity for the ring. We conclude that this finding for RG1 vs. a fairly steady gradient for more metal-poor regions further from the centre for disk galaxies and a variety of gradients for PRGs, argues in favour interpreting the extended feature in RG1 as indeed a ring, and not part of a disk. Another argument in favour of considering RG1 a PRG and not a disky spiral is the character of the kinematics of the gas vs. stars in the central region. Not only have we observed counter-rotation of these two components, but the motions there occur on different planes. This is different for spiral galaxies that show counter-rotation; the motions there take place in the same plane and while counter-rotation is seen along the major axis, only a zero velocity gradient is seen along the minor axis (e.g., Sil'chenko et al. 2009). We summarize the arguments for or against these two possibilities for the nature of RG1 in the Truth Table~\\ref{t:truth}. The option of a PRG has more + signs than the other option thus, on the balance of probabilities, we accept it over the alternative of a barred spiral with tight spiral arms and a strong disk warp. Therefore RG1 could be a rare case of a galaxy where the polar ring shows only a small inclination with respect to the equatorial plane of the CB. \\begin{table} \\caption{Two options to explain RG1} \\label{t:truth} \\begin{tabular}{ccc} \\\\ \\hline Property & Disk & PRG \\\\ \\hline Elongated CB connected with ring & + & -- \\\\ Relative brightness ring/CB & -- & + \\\\ Colours CB/ring & + & + \\\\ Ring metal abundance & -- & + \\\\ Ring radial brightness distribution & -- & + \\\\ Central gas (H$\\beta$) vs. stars kinematics & -- & + \\\\ CB velocity dispersion & -- & + \\\\ Tully-Fisher & + & -- \\\\ Cluster neighbourhood & + & -- \\\\ \\hline \\end{tabular} \\end{table}" }, "0910/0910.0540_arXiv.txt": { "abstract": "Using the IRAC images from the {\\it Spitzer c2d} program, we have made a survey of mid-infrared outflows in the $\\rho$ Ophiuchi molecular cloud. Extended objects that have prominent emission in IRAC channel 2 (4.5 \\micron) compared to IRAC channel 1 (3.6 \\micron) and stand out as green objects in the three-color images (3.6 \\micron \\ in blue, 4.5 \\micron \\ in green, 8.0 \\micron \\ in red) are identified as mid-infrared outflows. As a result, we detected 13 new outflows in the $\\rho$ Ophiuchi molecular cloud that have not been previously observed in optical or near-infrared. In addition, at the positions of previously observed HH objects or near-infrared emission, we detected 31 mid-infrared outflows, among which seven correspond to previously observed HH objects and 30 to near-infrared emission. Most of the mid-infrared outflows detected in the $\\rho$ Ophiuchi cloud are concentrated in the L1688 dense core region. In combination with the survey results for Young Stellar Objects (YSOs) and millimeter and sub-millimeter sources, the distribution of mid-infrared outflows in the $\\rho$ Ophiuchi molecular complex hints a propagation of star formation in the cloud in the direction from the northwest to the southeast as suggested by previous studies of the region. ", "introduction": "Mass outflows play an essential role in the process of star formation and have been found to be ubiquitous in various stages of star formation \\citep{arc07,bal07}. The specific angular momentum of a star-forming molecular core is about four orders of magnitude times higher than that of a typical T Tauri star \\citep{bod95}. One way to transfer the excessive angular momentum from the star-forming cores so that the circumstellar material can be accreted onto the central stars is through mass outflows \\citep{sha07}. Mass outflows from YSOs have been observed at different wavelengths. In the visual, high velocity jets and HH objects with a typical velocity of 100 - 300 km s$^{-1}$ trace material ejected by the star or shocked surrounding medium \\citep{rei01}. Near-infrared molecular hydrogen emission features, which usually have velocities of several tens km s$^{-1}$, trace H$_2$ gas in the jets or in the entrained surrounding medium \\citep{eis00}. In the millimeter wavelength, high velocity CO gas with velocity in the range from a few to ten km s$^{-1}$ have been observed \\citep{wu04}. The high velocity CO gas probes the swept or entrained medium. The different velocity ranges and excitation conditions for HH objects, near-infrared molecular hydrogen emission, and CO outflows indicate the transfer of energy and momentum from the high velocity material to the surrounding medium. Shocked gas in outflows has abundant atomic and molecular hydrogen line emission in the mid-infrared. Hydrodynamic simulations of outflows from YSOs by \\citet{sr05} have shown that outflows have strong H$_2$ line emission in all the IRAC four bands. In fact, mid-infrared observations with IRAC aboard of {\\it Spitzer} have discovered a lot of new outflows \\citep{mer08} and plenty new features of previously known outflows \\citep{nor04,smith06,tei08}, showing that {\\it Spitzer} IRAC imaging is a powerful tool in the survey of outflows, particular for those that are deeply embedded in clouds due to the much less extinction in the mid-infrared compared to the optical and near-infrared wavelengths. The {\\it Spitzer} legacy program {\\it c2d} performed a complete imaging of five nearby star-forming regions with the IRAC and MIPS instruments \\citep{eva03,eva09}, which provides a valuable data set to search for mid-infrared outflows in these regions. In this paper we present the results of a mid-infrared outflow survey of the $\\rho$ Ophiuchi molecular cloud based on the {\\it c2d} archive images. The $\\rho$ Ophiuchi molecular cloud is one of the closest star-forming regions at a distance of about 130 pc \\citep{wil08}. It is intermediate in star formation activity compared to the isolated star formation in the Taurus cloud and the rich clusters in the Orion cloud. The large scale structure of the $\\rho$ Ophiuchi cloud has been revealed by extensive molecular $^{13}$CO line mapping \\citep{lor89a,lor89b}. Millimeter and submillimeter surveys of the central part of the cloud have revealed the clumpy structure of the cloud and have found about 55 starless cores with masses in the range of 0.02 to 6.3 M$_\\sun$ \\citep{motte98,johnstone00,you06}. The cloud has been surveyed for young stars at wavelengths from X-ray \\citep{montmerle83,imanishi01,ozawa05}, visual \\citep{wilking87,wilking05}, near-infrared \\citep{gre92,allen02}, mid- and far-infrared \\citep{you86,wilking01,pad08}, to millimeter and submillimeter \\citep{sta06,andrews07}. Based on these surveys, \\citet{wil08} compiled a list of 316 young stars in L1688, the main cloud of the $\\rho$ Ophiuchi complex. About 46 HH objects \\citep{wilking97,wu02,phe04}, 119 H$_2$ emission features \\citep{gro01,gom03,kha04}, 16 high velocity CO outflows \\citep{bontemps96,bussmann07} have been observed in the $\\rho$ Ophiuchi complex. For a summary of the $\\rho$ Ophiuchi complex we refer to the recent review by \\citet{wil08} ", "conclusions": "" }, "0910/0910.5223_arXiv.txt": { "abstract": "We present Hubble Space Telescope optical coronagraphic polarization imaging observations of the dusty debris disk HD 61005. The scattered light intensity image and polarization structure reveal a highly inclined disk with a clear asymmetric, swept back component, suggestive of significant interaction with the ambient interstellar medium. The combination of our new data with the published 1.1 $\\mu$m discovery image shows that the grains are blue scattering with no strong color gradient as a function of radius, implying predominantly sub-micron sized grains. We investigate possible explanations that could account for the observed swept back, asymmetric morphology. Previous work has suggested that HD 61005 may be interacting with a cold, unusually dense interstellar cloud. However, limits on the intervening interstellar gas column density from an optical spectrum of HD 61005 in the Na I D lines render this possibility unlikely. Instead, HD 61005 may be embedded in a more typical warm, low-density cloud that introduces secular perturbations to dust grain orbits. This mechanism can significantly distort the ensemble disk structure within a typical cloud crossing time. For a counterintuitive relative flow direction---parallel to the disk midplane---we find that the structures generated by these distortions can very roughly approximate the HD 61005 morphology. Future observational studies constraining the direction of the relative interstellar medium flow will thus provide an important constraint for future modeling. Independent of the interpretation for HD 61005, we expect that interstellar gas drag likely plays a role in producing asymmetries observed in other debris disk systems, such as HD 15115 and $\\delta$ Velorum. ", "introduction": "Nearly two dozen dusty debris disks surrounding nearby stars have now been spatially resolved at one or more wavelengths. Many of these systems show clear similarities. For example, the radial architecture of several debris disks can be understood in terms of a unified model of steady-state dust production via collisions in a parent planetesimal belt (e.g., Strubbe \\& Chiang 2006). However, while the observed structure of many systems is ring-like \\citep{Kalas06,Wyatt08}, most disks show substructure such as clumps, warps, offsets, and brightness asymmetries not explained in traditional steady-state collisional grinding models. These unexpected features have triggered a great deal of recent theoretical work. The effects of massive planetesimal collisions, sandblasting by interstellar grains, close stellar flybys, dust avalanches, and secular and resonant perturbations by exoplanets have all been invoked to explain the observations (e.g., Moro-Martin et al. 2007, and references therein). However, as many of these theories produce similar structures, further observational constraints are needed to better understand the key forces affecting disk structure and the circumstances in which they apply. At a heliocentric distance of 34.5 pc \\citep{Perryman97}, the debris disk surrounding HD 61005 (SpT: G8 V; Gray et al. 2006), is a promising target for advancing our understanding in this area. The significant infrared excess for this source ($L_{\\rm IR}/L_{*} = 2 \\times 10^{-3}$) was recently discovered as part of the Spitzer FEPS survey \\citep{Carpenter09}, indicating 60 K blackbody-emitting grains $\\gtrsim$ 16 AU from the star. Follow-up Hubble Space Telescope (HST) coronagraphic imaging observations with the Near Infrared Camera and Multi-Object Spectrometer (NICMOS; HST/GO program 10527; D. Hines, PI) resolved the source (Hines et al. 2007, hereafter H07), revealing an unprecedented swept, asymmetric morphology, suggestive of significant interaction with the interstellar medium (ISM). H07 suggested that this system could be a highly inclined debris disk, undergoing ram pressure stripping by the ambient ISM. However, this interpretation requires an unusually high interstellar density for the low-density Local Bubble in which HD 61005 resides. Furthermore, the single wavelength intensity image was insufficient to provide strong constraints on the dominant size of the scattering grains and the overall scattering geometry. To further quantify the physical properties of grains seen in scattered light and the overall geometry of the system, we obtained optical coronagraphic polarimetry imaging observations of HD 61005 with the Advanced Camera for Surveys (ACS) onboard HST. As demonstrated by \\citet{Graham07} for the case of AU Mic, polarization observations in scattered light are invaluable for breaking degeneracies between grain scattering properties and their spatial distribution. Furthermore, the ACS data represent a factor of two improvement in angular resolution compared to the 1.1 $\\mu$m discovery observations. In addition to these new imaging data, we also obtained a high resolution optical spectrum to characterize ambient interstellar gas surrounding this system. In \\S 2, we describe the steps taken in observing and reducing these data. In \\S 3, we discuss the results of these observations, their consequences for the system scattering geometry, and the additional constraints they provide when combined with the 1.1 $\\mu$m NICMOS image. In \\S 4, we explore whether interactions with ambient interstellar gas can plausibly explain the observed swept, asymmetric morphology in this system. We discuss the implications for these potential explanations in \\S 5 and summarize our findings in \\S 6. ", "conclusions": "\\subsection{Interstellar Gas and the HD 61005 Morphology} The previous section explored whether disk/gas interaction can plausibly explain the unusual HD 61005 morphology. Of the four scenarios considered, three are implausible, given the limits on the ambient interstellar gas density imposed by our optical spectrum. The fourth scenario, secular perturbations from low density gas, is an attractive alternative, as this mechanism can significantly distort grain orbits well within a cloud crossing time. Furthermore, the densities required by this scenario are typical of local interstellar clouds, which occupy up to $\\sim20$\\% of the local ISM. Nevertheless, our preliminary modeling of this effect (Figures \\ref{vr} $-$ \\ref {vz}) can only produce disk morphologies in very rough agreement with the observations, suggesting that either additional physics needs to be incorporated into the current models, or that an altogether distinct physical mechanism is at work. Indeed, the current models are simplistic, and their applicability is limited by several key assumptions: \\begin{enumerate} \\item{{\\it Astrosphere Sizes:} As discussed by \\citet{Scherer00}, the toy models presented in Figure \\ref{vr} $-$ \\ref {vz} require that the disturbed grains be inside the astrospheric termination shock, such that the interstellar gas density and velocity can be approximated as constant. For the case of HD 61005, the termination shock distance is unknown. Furthermore, as HD 61005 is farther away than any star for which a direct astrospheric detection has been made \\citep{Wood04}, the astrosphere size may be difficult to constrain observationally. In general, termination shock distances vary greatly, depending on the ambient ISM and stellar wind conditions (e.g., densities, temperatures, velocities, stellar activity). For the case of the Sun only, the hydrodynamic models of \\citet{Muller06} show that the termination shock distance could easily vary between $\\sim 10$ AU and $500$ AU. Observational astrosphere measurements of solar-type stars are consistent with these predictions\\footnote{ On a broader scale, no astrosphere detections have been made for stars earlier than G-type. As a result, the typical effect of astrospheres on IS gas drag in debris disks surrounding A-type stars is difficult to reliably assess.} (Mann et al. 2006, and references therein). } \\item{{\\it Initial Conditions:} As noted in \\S 4.2.2 and Appendix B, our adopted initial orbital elements for the HD 61005 disk grains prior to the interstellar cloud interaction are highly uncertain, given a lack of information for the grain properties and underlying planetesimal population that collisionally replenishes the observed dust disk. Future long-wavelength observations sensitive to larger grains may be able to place tighter constraints on the distribution of sub-micron grains prior to the interstellar cloud interaction, as the distribution of large grains would likely reflect that of the parent bodies for the sub-micron population. Furthermore, larger grains would not be significantly affected by interstellar gas drag on the same timescale as the sub-micron size grains traced in these observations. The numerical techniques employed here and in \\citet{Scherer00} can be easily revised to accommodate an arbitrary initial disk architecture, provided the orbits are not highly eccentric, such that averaging over one orbit and applying Gauss' method is invalid.} \\item{{\\it Internal Disk Collisions:} In Appendix B, we estimate to order-of-magnitude that the collision time for submicron grains at $\\sim$70 AU is $\\sim$5000 yr. This collision time is somewhat longer than the timescale over which our model relaxes to a steady state --- essentially the time for gas drag to unbind a grain --- given in Appendix B as $\\sim$3000 yr. That the times are comparable supports the assumption of our models that each grain removed by gas drag is collisionally replenished. At the same time, the comparison of timescales underscores a shortcoming of our model --- that removal of grains by collisions is ignored. In reality, submicron grains should be removed from the system not only by gas drag, but also by collisions, in roughly equal proportions. We defer to future work a comprehensive study that includes removal by collisions via a collisional cascade.} \\item{{\\it Planetary Configurations:} Finally, as illustrated by \\citet{Scherer00} (e.g., his Figure 2), the incorporation of planetary orbits can appreciably change the perturbed orbital elements from the case in which only IS gas drag is considered. This caveat is especially important for massive grains, or grains in close proximity to planetary orbits. As a result, the models presented here should be treated with some caution if applied to typical planetary system scales ($\\lesssim 50$ AU; e.g., Kenyon \\& Bromley 2004). This cautionary point may be particularly important for the case of HD 61005, as the origin of the brightness asymmetry between the northeast and southwest disk lobes (\\S 3.1.1) is unknown. The agreement between the northeast and southwest deflected component position angles (\\S 3.1.3 and Figures \\ref{acs_image_log} $-$ \\ref{acs_image_lin}) suggests this asymmetry may originate from a physical mechanism entirely distinct from ISM interaction. If the asymmetry is due to a massive perturber, the disk morphologies produced in Figures \\ref{vr} $-$ \\ref{vz} are likely to be inapplicable. Resolved long wavelength observations sensitive to massive grains are needed to further explore this possibility.} \\end{enumerate} In addition to the above uncertainties, a remaining ambiguity important for future IS gas drag modeling is the velocity of the putative cloud responsible for the HD 61005 morphology. H07 noted that the star's tangential space motion is perpendicular to the disk midplane, in agreement with the relative flow vector suggested by initial inspection of the observed images. However, while assigning the relative flow direction to the star's tangential velocity is appealing, the three scenarios explored in \\S 4 which assume this flow direction were found to be untenable. Furthermore, velocities of local warm clouds can be comparable to the observed space motion of HD 61005 (\\S 4.2.2). As a result, the star's tangential motion is not a reliable indicator of the cloud-star relative velocity. Future spectroscopic observations may be able to detect the cloud directly (e.g., HST/GO Program 11674; H. Maness, PI), thereby providing key constraints on the ambient ISM density and velocity. Such observations will greatly inform future modeling, as the preliminary interstellar gas drag models presented here suggest a counterintuitive relative motion parallel to the disk midplane, rather than perpendicular to it. Though HD 61005 is difficult to assign to any known interstellar clouds (\\S 4.2.2), its galactic coordinates could plausibly associate it with either the G cloud or the Blue cloud. The space velocities of both of these clouds suggest a relative motion dominated by the radial velocity component. Therefore if HD 61005 is associated with either of these clouds, the relative motion is inconsistent with all models posited in \\S 4. \\subsection{General Applicability of Interstellar Gas Drag} The normalcy of the interstellar densities, velocities, and cloud sizes required by the secular perturbation model in \\S 4.2.2 suggests that IS gas drag can be important beyond HD 61005 in shaping debris disk morphologies. Taking the simple models of \\S 4.2.2 and Appendix B at face value, several of the general morphological features produced in Figures \\ref{vr} $-$ \\ref{vz} are consistent with observed disk structures. For example, the extreme brightness asymmetry in HD 15115 \\citep{Kalas07} may potentially result from interstellar gas drag, though a range of alternative explanations could explain this system as well (e.g., see list in \\S 1). The bow structures seen in some of the face-on models in Figure \\ref{vr} are also reminiscent of the mid-infrared morphology observed around the A star, $\\delta$ Velorum, which was recently modeled as a purely interstellar dust phenomenon \\citep{Gaspar08}. Finally, the middle panels in Figure \\ref{vrvz} show that warps, similar to that seen in $\\beta$ Pic (e.g., Mouillet et al. 1997), can in principle be produced for a relatively wide range of flow directions. However, while IS gas drag can in principle produce commonly observed disk features, the rate at which gas drag is expected to affect the observations remains unclear. Beyond the uncertainties in the model physics described in \\S 5.1, the characteristics of warm, low density clouds are currently uncertain, as detailed knowledge of them is limited to clouds residing predominantly within 15 pc of the Sun \\citep{Redfield08}. As a result, our understanding of typical cloud sizes, shapes, and total volumetric filling factor remains rudimentary. A key finding in this area, however, is that a significant fraction of nearby warm clouds appear to exhibit filamentary morphologies, which would limit the average interaction time between a given disk and cloud, likely reducing the rate at which IS gas perturbations produce an observable effect. This concern is particularly important for the case of disks surrounding early-type stars, as grains traced in scattered light tend to be larger in disks surrounding A-type stars than in their later type counterparts, owing to the larger radiation pressure blowout size. As such, the scattered-light morphologies for A-star disks require correspondingly longer cloud-disk interaction times to be noticeably affected. The timescale for a given grain to become unbound under IS gas drag increases approximately as the square root of the grain size \\citep{Scherer00}. \\subsection{Interstellar Grains and the HD 61005 Morphology} Finally, we note that all models posited in \\S 4 consider only the role of interstellar gas, ignoring the potential effects of interstellar grains. \\citet{Artymowicz97} investigated IS sandblasting of debris disks surrounding A stars and found that sandblasting has a negligible effect on the observed structure, as radiation pressure blows out most incoming interstellar grains before they are allowed to intersect the disk. However, under this framework, only grains with $\\beta \\geq 1$ are ejected. Thus Figure \\ref{radpr} shows that radiation pressure does not protect the HD 61005 disk, as it does in A-stars. Nevertheless, even if radiation pressure does not protect the disk against sandblasting, the stellar wind might, as only large interstellar grains with sizes greater than a few $\\times 0.1 \\mu$m are allowed to enter astrospheres freely without deflection \\citep{Linde00}. Thus it is likely that interstellar sandblasting can only plausibly compete with interstellar gas drag if the astrosphere is smaller than or comparable to the observed debris disk size \\citep{Mann06}. The size of the HD 61005 astrosphere is unconstrained by present observations. In general, observations and models of astrospheres surrounding solar-type stars show sizes in the range $\\sim 10-10^3$ AU, depending on the ambient ISM and stellar wind conditions (see discussion in \\S 5.1). Thus with a characteristic disk size of $\\lesssim 70$ AU, it is not clear whether typical interstellar grains can intersect the HD 61005 disk. Detailed modeling of sandblasting is outside the scope of this paper. However, future theoretical work should investigate the effects of sandblasting on debris disks surrounding solar-type stars. Calculations of the ISM density required for sandblasting to eject an observable flux of grains, the disk morphologies produced in this case, and the timescale for which sandblasting can be sustained would significantly aid in differentiating between this explanation and the gas drag models presented here." }, "0910/0910.5824_arXiv.txt": { "abstract": "We have used the zCOSMOS-bright 10k sample to identify 3244 {\\em Spitzer/MIPS} \\tfm-selected galaxies with $0.06< S_{24 \\, \\rm \\mu m}\\lsim 0.50 \\, \\rm mJy$ and $I_{\\rm AB}<22.5$, over 1.5~deg$^2$ of the COSMOS field, and studied different spectral properties, depending on redshift. At $0.21$, the IR background is mainly dominated by luminous and ultra-luminous IR galaxies (LIRGs and ULIRGs, with total IR [$5-1000 \\, \\rm \\mu m$] luminosities $10^{11} 10^{12} \\, \\rm L_\\odot$, respectively). However, these populations are progressively less important at lower redshifts, becoming rare by the present epoch. The most typical IR sources at $z \\lsim 0.7$ are IR normal galaxies, i.e. those with total IR luminosities $L_{\\rm TIR}< 10^{11} \\, \\rm L_\\odot$. Their study is fundamental to clarify the production of the bulk of the IR background in the second half of cosmic time. Our knowledge of IR galaxies at low redshifts was first possible thanks to studies conducted with the {\\em Infrared Astronomical Satellite (IRAS)} and the {\\em Infrared Space Observatory (ISO)} (see e.g. Aussel et al.~1999; Genzel \\& Cesarsky~2000, for a review), which revealed that both star formation and nuclear activity were associated with the dust emission. More recently, with the advent of the {\\em Spitzer Space Telescope} (Werner et al.~2004) and the {\\em Akari Telescope} (Matsuhara et al.~2006), a much deeper insight into the nature of the $z<1$ IR galaxy population has been possible. We have now characterised their luminosity evolution (e.g. Le Floc'h et al.~2005; Caputi et al.~2007; Magnelli et al.~2009), stellar masses (e.g. Caputi et al. 2006a,b), and spectral energy distributions (SEDs; e.g. Rowan-Robinson et al.~2005; Rocca-Volmerange et al.~2007; Takagi et al.~2007; Bavouzet et al.~2008; Kartaltepe et al.~2009; Symeonidis et al.~2009), among others (see Soifer et al.~2008, for a review). Spectroscopic surveys at different wavelengths can add very valuable information that help understand the nature of the $z<1$ IR sources. In addition to providing accurate redshift determinations, spectroscopic data allow us to characterise other aspects of IR galaxies, such as their extinction properties, metallicities and general chemical composition. The initial optical/near-IR spectroscopic studies of IR galaxies at $z<1$ targeted {\\em IRAS} and {\\em ISO} sources (e.g. Kim et al.~1995; Veilleux et al.~1995, 1999; Franceschini et al.~2003; Flores et al.~2004; Arribas et al.~2008; Garc\\'{i}a-Mar\\'{i}n et al.~2009). In the {\\em Spitzer} era, and in the light of larger spectroscopic campaigns, more systematic studies of IR galaxies and their environments have been possible (e.g. Choi et al.~2006; Papovich et al.~2006; Caputi et al.~2008 [C08 hereafter], 2009; Desai et al.~2008). In particular, optical spectra are useful to disentangle the power source of the IR emission, which generally occurs when the galaxy dust re-processes the energy provided by the ultra-violet (UV) photons of young stars or active nuclei. However, in some cases, the observed IR emission might not be associated with any on-going stellar or nuclear activity, but be rather produced by dust heated by old stars (e.g. Gordon et al.~2000; Kong et al.~2004; Cortese et al.~2006; but also see Young et al.~2009), or even direct stellar emission. The importance of these mechanisms within the IR galaxy population is not yet well established. The identification of `passive' IR sources is important to quantify the degree of contamination in the IR-derived cosmic star formation rate densities. Spectroscopic data can be very helpful to elucidate the presence of star formation in sources that are presumably dominated by old stellar populations. Spectral features also carry the imprints of the galaxy star formation history. Thus, for IR galaxies, optical spectra can be used to investigate at what stage of the galaxy history the IR phase is produced. Locally, the rare LIRGs and ULIRGs are known to be mainly the consequence of galaxy mergers that induce bursts of star formation and nuclear activity (e.g. Armus et al.~2007). Instead, more typical IR galaxies at $z=0$ are usually less disturbed, and appear to form their stars in a more regular way. There are several clues that suggest that, in the past, when LIRGs and ULIRGs were more common, both the burst-like and quiescent modes of star formation could also lead to the IR phase (e.g. Elbaz et al.~2007; Daddi et al.~2008). In spite of these different pieces of evidence, it is not yet clearly established whether there exists a transition in the possible modes of star formation between galaxies of different IR luminosities. One of the main aims of this work is to study such a transition in IR galaxies at $0.6 5 \\, \\rm \\AA$, which indicates that the vast majority of IR sources at these redshifts are associated with significant on-going star formation. In most of these sources, the relation between the gas obscuration, i.e. $L(\\rm H\\alpha)/L(\\rm H\\beta)$, and IR/UV attenuation can be reasonably explained with different reddening laws of common use in the literature. However, for up to 16\\% of our sample at $0.2 5 \\, \\rm \\AA$ are in the blue ($B-i<1.3$) cloud, and around 70\\% of those with \\hb EW $< 5 \\, \\rm \\AA$ are in the `green valley' or redder ($B-i \\geq 1.3$). $\\bullet$ The location in the BPT diagram of our galaxies with \\ha EW $> 5 \\, \\rm \\AA$ suggests that $\\sim 22\\%$ of them are composite systems. However, in none of these galaxies can the presence of an active nucleus be confirmed with the X-ray or {\\em IRAC} data available in the COSMOS field. This fact suggests that, although an AGN component might be present in these galaxies, in any case is not dominant, and that aperture effects could also influence the fraction of sources identified as composites. $\\bullet$ We identified only 5 galaxies in our sample at $0.20.5$. We used the zCOSMOS spectra to determine oxygen abundances, measure the $\\rm D_n(4000)$ parameter and the average $\\rm H_\\delta$ EWs. Our results are the following: $\\bullet$ 79\\% of our galaxies with \\hb EW $> 5 \\, \\rm \\AA$ at $0.5